You are on page 1of 11

Fuel Processing Technology 176 (2018) 296–306

Contents lists available at ScienceDirect

Fuel Processing Technology


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

Research article

Particulate and gaseous emissions from charcoal combustion in barbecue T


grills

Vicente E.D.a, Vicente A.a, Evtyugina M.a, Carvalho R.a,b, Tarelho L.A.C.a, Oduber F.I.c, Alves C.a,
a
Centre for Environmental and Marine Studies, Department of Environment and Planning, University of Aveiro, 3810-193 Aveiro, Portugal
b
Department of Applied Physics and Electronics, Umeå University, 901 87 Umeå, Sweden
c
Department of Physics, IMARENAB, University of León, 24071 León, Spain

A R T I C LE I N FO A B S T R A C T

Keywords: The use of charcoal for cooking and heating can be a major source of air pollution and lead to a wide range of
Barbecue grill health outcomes. The aim of this study was to experimentally quantify and characterise the gaseous and par-
Charcoal ticulate matter (PM2.5) emissions from charcoal combustion in a typical brick barbecue grill. The gaseous
Emissions emission factors were 219 ± 44.8 g kg−1 for carbon monoxide (CO), 3.01 ± 0.698 g kg−1 for nitrogen oxides
Chemical composition
(NOx expressed as NO2), and 4.33 ± 1.53 gC kg−1 for total organic carbon (TOC). Particle emissions
PM2.5
(7.38 ± 0.353 g kg−1 of dry charcoal burned) were of the same order of magnitude as those from traditional
Organic markers
residential wood burning appliances. About 50% of the PM2.5 emitted had a carbonaceous nature while water
soluble ions accounted, on average, for 17% of the particulate mass. Alkanes (C11–C16 and C23), hopanes,
steranes and alkyl-PAHs accounted for small mass fractions of PM2.5. Phenolic compounds and saccharides
represented the major particle-bond organic constituents. The high proportion of either resin acids or syringyl
and vanillyl compounds is consistent with emissions from charred coniferous wood. The ratios between anhy-
drosugars for charcoal are much lower than the values reported for lignite combustion, but overlap those from
other biomass burning sources.

1. Introduction charcoal has a carbon footprint 2.9 times higher [3].


Charcoal production in Africa accounted for 61% of global pro-
In 2014, about 53 million tonnes of wood charcoal were produced duction in 2014. In the Latin America and Caribbean, charcoal pro-
worldwide [1]. Charcoal is a product of thermochemical conversion of duction grew from 2010 to 2014, reaching 10 million tonnes [1]. Sev-
biomass by pyrolysis and has advantages as fuel when compared to the eral studies focused on the emissions from charcoal production have
original feedstock (biomass) such as the higher heating value and easy been performed [5,7–10]. In 2009, the greenhouse gas emissions
storage [2,3]. It has also an advantage in comparison with other re- arising from charcoal production in tropical ecosystems were estimated
newable fuels: the cheaper production. In contrast with mineral coals, to be around 71.2 million tonnes of CO2 and 1.3 million tonnes of CH4
charcoal has relatively low content of sulphur or mercury, which is a [6]. Charcoal is produced from wood pyrolysis in kilns and the process
benefit from the emissions point of view [2]. In the developing coun- may take up to a few weeks [2]. The efficiency of traditional methods
tries, charcoal is still an important cooking fuel [4]. Despite the spatial for charcoal production is about 10% to 22%. Retort kilns have been
and temporal changes in fuel consumption patterns, charcoal remains a claimed to increase energy efficiency and to decrease air pollution ef-
popular cooking fuel in the developed world since it produces food with fects [5]. Sparrevik et al. [7] confirmed that retort kilns lead to sig-
unique flavour and texture. In fact, charcoal-grilling is extensively used nificantly lower emissions of incomplete combustion products. How-
by households and restaurants [3,5]. Johnson [3] estimated the char- ever, the efficiency from these kilns is lowered due to the wood
coal grilling footprint to be 6.7 kg CO2e per grill session which is, ac- consumption for start-up. In addition to the production process, char-
cording to the author, similar to the carbon footprint obtained for coal combustion is a source of airborne pollution. Intense outdoor
driving an average car for approximately 35 km. Charcoal production is barbecue cooking during a big festival event was reported to increase
the dominant process in the carbon footprint (45%), as well as charcoal the ambient PM10 levels by approximately 5% [11]. Charcoal burning
combustion (40%). Compared to liquefied petroleum gas (LPG) grilling, leads to the emission of a wide range of pollutants. Alves et al. [12]


Corresponding author.
E-mail address: celia.alves@ua.pt (C. Alves).

https://doi.org/10.1016/j.fuproc.2018.03.004
Received 8 January 2018; Received in revised form 4 March 2018; Accepted 4 March 2018
0378-3820/ © 2018 Elsevier B.V. All rights reserved.
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

collected PM2.5 samples from the exhaust stacks of restaurants with These, in turn, are key aspects concerning the emissions generated by
different cooking styles including a charcoal grilled chicken restaurant. charcoal combustion (e.g., [18]). For these reasons, it is advantageous
The PM2.5 concentrations reported by the authors were in the range of to obtain source-specific emission profiles taking into account the fuel
26 to 127 mg m−3. For the same sites, the VOC concentrations were also specificities.
assessed [13]. The authors observed that chlorinated VOCs were only The aim of this study was to experimentally quantify and char-
detected in samples from the charcoal grilled chicken restaurant. Ben- acterise the particulate and gaseous emissions from charcoal combus-
zene was the compound with the highest emission rate (201 kg year−1). tion in a typical Portuguese barbecue grill. In doing so, the data pre-
In indoor environments, such as restaurants, personal exposure to sented in the current study will improve emission inventories, which
toxic pollutants from charcoal burning can be significantly high, posing are indispensable to establish environmental measures to prevent air
a health risk to the customers and employees. Lee et al. [14] assessed pollution. Moreover, new databases of speciated emission profiles will
the indoor air quality at four different types of restaurants. The authors contribute to more accurate source apportionment results when ap-
found the highest CO (15,100 μg m−3) and particulate concentrations plying receptor models.
(1442 and 1167 μg m−3 PM10 and PM2.5, respectively) at the barbecue
style restaurant. High PM10 (15,074 μg m−3) and PM2.5
2. Methodology
(13,700 μg m−3) concentrations were also found in a Korean-style
barbecue restaurant in Seoul [15]. Moreover, barbecue cooking can
2.1. Combustion infrastructure, fuel and experimental procedure
generate considerable amounts of benzene, toluene, methylene chloride
and chloroform [14]. Chinese traditional charbroiling was found to be a
The experimental setup of the present study replicates the
source of particle-bound polycyclic aromatic hydrocarbons (PAHs) and
Portuguese charcoal barbecue grills, which have side and back walls
fatty acids indoors. The latter accounted for over 90% of all identified
made of refractory brick and a chimney hood. The experimental setup
organic compounds. However, PAHs had a major contribution to the
included a grate where the charcoal was burned and which was placed
particulate mass [16]. Taner et al. [17] carried out a study at 14 non-
over a weight sensor (DSEUROPE Model 535QD-A5), allowing the
smoking restaurants in Turkey. The authors concluded that fine parti-
continuous monitoring of the fuel mass during the combustion cycles.
cles associated with charcoal cooking represented a significant source
The combustion flue gas flow rate across the chimney was calculated by
of indoor pollution. Charcoal combustion played an important role in
monitoring the gas velocity with a Pitot tube connected to a differential
the PM trace element content and the most abundant elements identi-
pressure transmitter (JUMO, Type 404304). A detailed description of
fied were As, Cr, Se, V, and Zn.
the experimental setup can be found elsewhere [27].
Pollutant emissions from charcoal burning are associated with the
The charcoal used in the combustion experiments was purchased
raw materials nature and the charcoal production process [18–21].
from a local supplier. The fuel properties were determined according to
Kabir et al. [22] investigated trace metal emissions from eleven types of
international CEN/TS standards and included moisture (CEN/TS
charcoal available in the Korean market (Korean and imported pro-
14774), ash content (CEN/TS 14775), volatile matter (CEN/TS 15148),
ducts). The combustion tests were carried out in an old type Korean
C, H, N, S (CEN/TS 15104) and calorific value (CEN/TS 14918)
burner. Although the trace metal concentrations varied among different
(Table 1). In order to replicate the householder's practices, the tests
charcoal types, Fe and Zn were consistently the most abundant metals.
were initiated by lightning small pieces of wood on the top of the batch
In an earlier study, Kabir et al. [19] reported aromatic volatile organic
of charcoal. A total of four cold start tests, initiated with the combustion
compounds and carbonyl emissions from the same charcoal types.
appliance at ambient temperature, were conducted. The experiments
Benzene and toluene were the most abundant VOCs, while for-
were made with batches of around 1 kg of charcoal fuel and lasted
maldehyde and acetaldehyde were the main carbonyls. The pre-
approximately 1.5 to 2 h. The combustion temperature was monitored
dominance of benzene over other aromatic compounds emitted from
using K-type thermocouples at three locations: under the fixed bed of
glowing charcoal was also reported by Olsson and Petersson [20].
fuel at the grate, at the central region of the combustion chamber
Charcoal burning is also a source of offensive odorants [21]. Barbecue
(0.25 m above the fixed bed of fuel) and at the chimney exit (2.6 m
charcoal combustion can be a significant source of trace metal emis-
above the fixed bed of fuel).
sions to the atmosphere. Susaya et al. [23] reported concentrations of
Cd, Co and Ni exceeding the inhalation minimum risk levels of the
United States Agency for Toxic Substances and Disease Registry for 2.2. Gas sampling and measurement techniques
chronic duration exposure. In their study, eleven types of charcoal were
burned in a combustion device built in a traditional Korean style with The combustion flue gas was sampled at the chimney through a
stainless-steel vent line. The major PM-bound metal elements were Zn heated (at 180 °C) sampling line, which conducted the gas to an online
and Pb. Although these and other studies provide remarkable data on
emissions from charcoal burning, only a few aimed at characterising the Table 1
Ultimate and proximate analysis of the charcoal used as fuel in the combustion
PM chemical composition, and these evaluations were mainly focused
experiments.
on the elemental content. Particles released during charcoal burning
contain a broad range of chemical species, ranging from elemental to Proximate analysis (wt%, as received)
organic and inorganic compounds. Very little is known about the par-
ticulate organic speciation. Due to their carcinogenic potential, only Moisture 4.06
Ash 11.0
PAHs have been addressed in a few studies [11,24,25]. In most re- Volatile matter 4.36
searches, only gaseous and particulate matter concentrations were Fixed carbon (by difference) 84.7
measured, not encompassing a more exhaustive characterisation of the
PM samples. Furthermore, the majority of the studies on this topic have Ultimate analysis (wt%, dry basis)

been focused on particular fuels from particular regions. A rather ex-


C 79.6
tensive amount of work has been performed to characterise the emis- H 2.5
sions from charcoal of the Asian market (e.g., [18,19,21–23,25]), and, N 1.03
to a lesser extent, few studies focused on charcoal fuels from Poland S < 0.01
O (by difference) 5.44
[20], Sweden [20,24] and United States [21,26]. The raw materials
used in charcoal production, as well as the production process condi- Lower heating value (MJ kg−1, dry basis) 29.3
tions, are very important regarding the final properties of the fuel [2].

297
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

Fourier Transform Infrared Gas analyser (FTIR Gasmet, CX4000) op- 1000
erating at 180 °C. This equipment enables real-time monitoring and Chimney
multicomponent measurement. Combustion chamber
Fuel bed
800
2.3. Particulate matter sampling and analysis

Temperature (ºC)
PM2.5 sampling covered the whole period of the combustion of a
charcoal batch. It was carried out after the full flue gas dilution (hood 600
dilution system) with atmospheric air in a tunnel installed downstream
the chimney. The contamination arising from the dilution air was
checked by background samples taken under the same conditions as the 400
combustion tests. The background levels have been subtracted from all
measurements in the dilution tunnel. The volumetric gas flow rate
throughout the tunnel was determined from the mean gas velocity
200
monitored by a Pitot tube and respective differential pressure sensor
(Testo AG 808) and a K-type thermocouple for temperature monitoring.
A stainless steel probe with an internal diameter of 11 mm was used in
order to ensure sampling under isokinetic conditions. The probe was 0

01:40:00
connected to a TCR TECORA (model 2.004.01) with a sampling flow

01:20:00
00:20:00

01:00:00
00:00:00

00:40:00
rate of 2.3 m3 h−1 (referred to normal temperature and pressure con-
ditions).
The PM samples for gravimetric and chemical analyses were col-
Time (hh:mm:ss)
lected on 47 mm diameter quartz fibre filters (Pallflex®) pre-baked at
500 °C for 6 h to remove organic contaminants. Prior to weighing, the Fig. 1. Temperature profile during the combustion of a batch of charcoal, inside the
filters were conditioned for around 24 h in a room with controlled combustion chamber and at the chimney exit.

humidity and temperature. The gravimetric quantification was per-


formed with a microbalance (RADWAG 5/2Y/F, accuracy of 1 μg) and down. The extracts were analysed by gas chromatography–mass spec-
obtained from the average of six measurements (relative standard de- trometry (GC–MS). Before injection, the compounds with hydroxylic
viation < 0.02%). and carboxylic groups were converted into the corresponding tri-
The PM2.5 organic (OC) and elemental carbon (EC) content was methylsilyl derivatives. N,O‑bis(trimethylsilyl)trifluoroacetamide
analysed by a thermal optical transmission technique. This method (BSTFA): trimethylchlorosilane (TMCS) 99:1 (Supelco 33149-U) was
relies on the CO2 quantification, by a non-dispersive infrared (NDIR) used as silylation reagent. A detailed description of the entire procedure
analyser, which is released from the volatilisation and oxidation of for the determination of organic compounds can be found in Alves et al.
different carbon fractions under controlled heating. A laser beam and a [28].
photodetector, measuring the filter light transmittance, allow separ-
ating the EC formed by OC pyrolysis from the one that was originally in
the sample. 3. Results and discussion
Water-soluble ions were analysed by ion chromatography. Two
filter punches of 9 mm were placed in a screwed cap vial and extracted 3.1. Operating conditions
with 6 mL of ultrapure Milli-Q water by ultrasonic agitation
(15 + 15 min). Extracts were filtered through a 13 mm PVDF syringe After starting the charcoal combustion process with kindling, the
filter with 0.2 μm pore size (Whatman™). The sample extracts were temperature rose rapidly during the initial 3 to 7 min of operation,
injected in an ion chromatograph (DIONEX, ICS-5000+ DC, USA) reaching values from 276 °C to 481 °C at the central region of the
which consists in a dual-system ion chromatograph-conductance de- combustion chamber. Once the maxima were attained, the temperature
tector (ICS-5000), an AS-DV 40 auto sampler and an eluent regenerator values gradually decreased until the full conversion of the fuel (Fig. 1).
system (Dual RFIC-EG, Dionex, USA) operating with electrolytic auto Considering the whole combustion cycle of a batch of charcoal, the
suppression recycle mode. For the analysis of anions, a DIONEX Ionpac mean temperature in the central region of the combustion chamber was
AS11-HC-4 μm analytical column (2 × 250 mm) with DIONEX Ionpac 123 ± 20.4 °C. The temperatures recorded at the chimney exit were
AG11-HC – 4 μm (2 × 50 mm) guard column and eluent 30 mM po- relatively low, and typically below 50 °C (Table 2). This is related to an
tassium hydroxide (KOH), were used. The analyses of cations were operation with a high air-to-fuel ratio due to the uncontrolled entrance
performed with a DIONEX Ionpac CS16 analytical column of combustion air in the chamber. The fuel consumption rate was es-
(3 × 250 mm) with DIONEX Ionpac CG16 (3 × 50 mm) guard column timated to be 0.748 ± 0.101 kg h−1 (Fig. 2). After the initial stage of
and eluent 30 mM methanesulphonic acid (MSA) This system enables charcoal ignition, visible flames were almost imperceptible, apart from
simultaneous injection in both cation and anion modules and de- very small flare-ups from the glowing embers, during the remaining
termination of all ions in a single run. period of combustion.
After carbon analysis and ion chromatography, the remaining por-
tions of the filters were combined in order to meet the detection limits
required by the speciated organic compounds. The pooled filter pieces
were extracted with dichloromethane (300 mL) for 24 h and then three Table 2
Temperature of the combustion flue gas at the fixed bed of fuel, at the central region of
times with methanol in an ultrasonic bath (25 mL for 10 min, each
the combustion chamber and at the chimney exit.
extraction). After each extraction, the extracts were combined and fil-
tered. The total organic extract was then concentrated by rotary eva- Chimney exit (°C) Combustion chamber (°C) Fuel bed (°C)
poration and, finally, separated into five different organic fractions of
Average 48.02 ± 2.62 123 ± 20.4 443 ± 69.4
increasing polarity by flash chromatography on a silica gel column
Minimum 18.18 ± 2.49 18.37 ± 2.2 17.8 ± 1.93
(activated at 150 °C during 3 h). After each elution, the different or- Maximum 84.34 ± 17.8 414 ± 93.8 767 ± 81.6
ganic fractions were vacuum concentrated and dried by nitrogen blow

298
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

1000 198 g kg−1 for the combustion of charcoal in traditional and improved
cookstoves used in different Asian countries. Their study also comprised
900 a comparison with the combustion of wood in the same stoves. The CO
EFs were slightly lower for wood-fired stoves (19 to 136 g kg−1). An-
dreae and Merlet [32] reported a CO EF of 200 ± 38 g kg−1 for dry
800 charcoal burning. Combustion of ten charcoal types from the Taiwan
market in a tubular high temperature furnace generated CO EFs ranging
Mass of fuel (g)

700 from 67.7 to 300 g kg−1 [18]. According to Huang et al. [18], the CO
emission during smouldering combustion is influenced by the charcoal
ash content. Higher ash content means that the charcoal surface is more
600
covered which can hinder the CO oxidation reaction. The ash content of
good-quality charcoal is often refereed as between 0.5 and 5% [2],
500 which is lower than the charcoal ash content used in the present study
(11.4% wt. dry basis).
The NOx EF (expressed as NO2) was 3006 ± 698 mg kg−1 of dry
400
charcoal burned. The average EF obtained of this study was similar to
the one reported by Andreae and Merlet [32] for charcoal burning.
300 However, the NOx emissions observed here are much higher than those
00:00:00

00:20:00

00:40:00

01:00:00

01:20:00

01:40:00
obtained in other studies on different kinds of charcoal burning
[18,31]. Huang et al. [18] reported NOx emissions in the range from
159 to 834 mg kg−1. Even lower NOx EFs (30 to 430 mg kg−1) were
Time (hh:mm:ss) recorded from mangrove charcoal combustion in different Asian cook-
Fig. 2. Evolution of the mass of fuel in the grate during the combustion of a charcoal stoves [31]. NO emissions from oil and natural gas combustion are
batch. mostly associated with thermal and prompt formation, whereas the
dominant mechanism of NO formation during combustion of biomass
fuels is from the fuel nitrogen content since the temperatures achieved
3.2. Gaseous emission factors
in the furnaces are usually too low for prompt and thermal formation
[33]. Usually the charcoal nitrogen content is lower than 0.6 wt% [2].
From the biofuel carbon content, 88–93% and 9–12% was emitted
In this study, charcoal has 1.03 wt% of nitrogen, which is slightly
in the form of CO2 and CO, respectively. The emission factors (EFs) of
higher when compared to the charcoal-fuels from the Taiwan market
CO and CO2 from charcoal combustion in the brick barbecue grill were,
[18].
on average, 219 ± 44.8 and 2619 ± 110 g kg−1 of dry charcoal
In the present study, the average TOC EF was 4.33 ± 1.53 gC kg−1
burned, respectively. The EF of CO obtained in this study was close to
of dry charcoal burned. The major average contributions to the re-
those reported in other works involving charcoal combustion (Table 3).
corded TOC emissions were from methane (73.2%) followed by ethane
CO is a product of incomplete combustion, generally associated with
(14.9%) and hexane (5.9%), while formaldehyde, ethane and propane
the smouldering phase. Charcoal combustion in inefficient appliances
accounted for minor fractions of 3.9, 1.1 and 1.1%, respectively
generates high levels of CO due to the low temperatures and reduced or
(Fig. 3). Methane, a relevant greenhouse gas, is an important compound
almost absence of a visible flame above the charcoal bed [26,29]. Jetter
released during biomass combustion [34]. Bhattacharya et al. [31] re-
et al. [30] investigated the combustion of six different fuels (wood,
ported similar CH4 emissions from wood and charcoal burning in
charcoal, pellets, corn cobs, rice hulls, and plant oil) at two fuel
cooking stoves. Emissions from a charcoal cooking fire, in Kaoma, were
moisture levels in 22 cookstoves, and reported CO emissions from
assessed with an OP-FTIR. The CH4 EF was 6880 mg kg−1 of dry
charcoal to be higher than the ones observed during wood combustion.
charcoal burned, which was 9 to 13 times higher than the emissions of
Bhattacharya et al. [31] reported CO EFs ranging from 34.2 to

Table 3
Comparison between gaseous EFs from this study and literature values.

Appliance Fuel CO (g kg−1) CO2 (g kg−1) NOx (mg kg−1) HCHO (mg kg−1) Reference

Tube furnace Eco-friendly charcoal 128 ± 22.1 776 ± 45.8 834 ± 491 8.0 ± 6.1 [18]
Charcoal briquettes 169 ± 20.5 836 ± 21.4 675 ± 93.6 520 ± 201
Charcoal briquettes 300 ± 31.3 987 ± 60.0 161 ± 98.6 9.8 ± 2.5
Sawdust briquette charcoal 215 ± 16.6 879 ± 96.3 462 ± 68.4 6.1 ± 0.4
Mangrove charcoal 109 ± 19.3 855 ± 92.1 277 ± 44.2 26.7 ± 11.4
Charcoal briquettes 67.7 ± 56.4 724 ± 43.2 414 ± 149.3 57.8 ± 10.4
Eco-friendly charcoal 180 ± 25.1 644 ± 53.1 159 ± 22.9 38.6 ± 14.8
Acacia charcoal 84.3 ± 1.0 1114 ± 83.9 824 ± 357 19.8 ± 2.9
Longan charcoal 108 ± 9.8 1097 ± 42.4 621 ± 209 15.1 ± 3.5
Binchōtan 76.6 ± 14.4 1225 ± 39.0 832 ± 296 16.2 ± 4.6
Cambodian traditional Mangrove charcoal 34.2 ± 3.2 2352 ± 13 70 ± 1 [31]
Thai-bucket cookstove 35.7 ± 3.9 2155 ± 26 30 ± 1
Chinese traditional 175 ± 11 2436 ± 54 300 ± 70
QB Philippine 198 ± 4 2276 ± 30 220 ± 30
Philippine traditional 155 ± 7 2567 ± 67 140 ± 20
Lao improved 134 ± 4 2451 ± 13 190 ± 10
Vietnamese improved 87.2 ± 5.7 2233 ± 61 300 ± 90
Malaysian improved 155 ± 16 2576 ± 59 430 ± 40
Bang sue stove 178 ± 17 2555 ± 70 420 ± 90
Brick barbecue grill Charcoal (Portuguese) 219 ± 44.8 2619 ± 110 3006 ± 698a 383 ± 90.3 This study

a
NOx expressed as NO2.

299
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

7 100

CH4
6 C2H6
80

gC kg of dry charcoal burned


C2H4
5 C3H8
C6H14
60
4

% TOC
3
40
-1

20
1

0 0
TOC CH4 C2H6 C2H4 C3H8 C6H14 HCHO TOC

Fig. 3. Hydrocarbon emission factors (expressed as gC kg−1 of dry charcoal burned) and average relative contribution to the measured TOC in combustion flue gases.

other organic chemical compounds (C2H6, C2H4, C2H2, HCHO) [35]. from wood combustion were 1.4 and 7.3 times higher than those re-
The CH4 EF obtained in the present study (4334 ± 1955 mg kg−1 of gistered for coal and charcoal combustion, respectively. The PM EF
dry charcoal burned) was higher than the one documented by Andreae reported by Oanh et al. [25] for charcoal burning in a Thai bucket stove
and Merlet [32] for charcoal burning (2700 mg kg−1, dry basis). (7 mg kg−1) was three orders of magnitude lower than the one obtained
Bhattacharya et al. [31] found CH4 EFs for wood (6000 to in this study. Huang et al. [18] found similar differences among the ten
10,000 mg kg−1) in the same range compared to charcoal (6700 to fuels tested. Charcoal briquettes generated a PM2.5 EF > 10 times
7800 mg kg−1). higher than that of other types of charcoal, 800 times that of the
SO2 is a pollutant arising from systems involving combustion of Binchōtan and 1.25 times higher than the one obtained in the present
sulphur containing fuels. However, the sulphur content in charcoal study for Portuguese charcoal (Fig. 5).
produced from natural wood is much lower when compared with that Alves et al. [38] obtained slightly higher PM2.5 emissions for wood
of coal or fuel oil. In fact, the elemental analysis of the charcoal used in combustion in a brick fireplace. Depending on the wood type burned,
the present study revealed a sulphur content under the quantification EFs were 1.25 to 2.56-fold higher than those of the present work. The
limit of the method (< 0.01 wt%), and thus the EF of SO2 is relatively same wood types were also burned in a woodstove by Alves et al. [38],
low (< 1.2 mg kg−1 of dry charcoal burned, Fig. 4). Likewise, the EFs of who reported PM EFs ranging from 6200 to 16,300 mg kg−1. Although
other acidic gases, HCl and HF, were relatively low (Fig. 4). As regards the average emission factor for wood burning was higher than the one
the alkaline gas NH3, the EF was (1.21 ± 0.235 mg kg−1 of dry char- found in the present study for charcoal combustion, the range of values
coal burned, Fig. 4) in the same order of magnitude of SO2. overlap. Thus, the PM emissions obtained can be considered compar-
able to those reported for wood burning in small-scale heating appli-
ances.
3.3. Particulate emission factors

The PM2.5 EF was 7380 ± 353 mg kg−1 of dry charcoal fuel 20000
burned. Oanh et al. [25] reported particulate emissions from three Huang et al. [18]
different Asian combustion systems and respective fuels. The authors 16000 This study
Alves et al. [38]
recorded the emissions from the open combustion of eucalypt wood 12000
chips, and charcoal (mangrove) in a Thai bucket stove and coal bri-
mg PM kg of fuel burned

8000
quettes in a Vietnamese cylindrical briquette stove. Particle emissions
4000
2.00
500
-1

1.60
400
1.20
g kg of dry charcoal burned

300
0.80
200

0.40 100
0.20
0
Binchotan
Eco-friendly charcoal 1

Charcoal briquettes 1

Charcoal briquettes 2

Eco-friendly charcoal 2

Portuguese woods-Fireplace

Portuguese woods-Woodstove
Charcoal briquettes 3
Mangrove charcoal

Acacia charcoal

Longan charcoal
Sawdust briquette charcoal
This study

0.15
-1

0.10

0.05

0.00
SO2 NH3 HCl HF

Fig. 4. Emission factors of gaseous compounds from charcoal combustion (g kg−1 of dry Fig. 5. Comparison between PM2.5 emission factors (mg kg−1) from different types of
charcoal burned). charcoal and wood combustion and the present study.

300
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

3.3.1. Water soluble ions 14 50


Other PAHs
Water soluble ions accounted for 17.2 ± 7.40 wt% of the PM2.5 Retene
12
mass. The inorganic mass fraction in PM2.5 was found to be inversely
40
correlated with the OC mass fraction (R2 = 0.938). The dominant ions

Other PAHs (µg g PM2.5)


10

Retene (µg g PM2.5)


in the charcoal combustion generated particles were potassium
(6.04 ± 3.04 wt%), sodium (4.52 ± 1.17 wt%) and chloride

-1
30
8

-1
(3.86 ± 1.91 wt%). Gonçalves et al. [36] reported high content of
these inorganic ions in particulate samples from the burning of wood
6
grown in coastal areas at high levels of salinity. Although detected, 20

calcium, nitrate, sulphate and fluoride, did not accounted for significant
4
percentages of the total mass of ions in the PM2.5. Some ions like bro-
10
mide, phosphate, nitrite, magnesium, ammonium and lithium were 2
below the detection limit of the method. Calvo et al. [37] reported
potassium and calcium as dominant water soluble cations for the 0 0
combustion of three common Southern and mid-European woods in two

Benzo[b]fluoranthene

Benzo[k]fluoranthene

Indeno[1,2,3-cd]pyrene
p-Terphenyl

Benzo[a]anthracene

Benzo[e]pyrene

Benzo[a]pyrene

Dibenzo[a,h]anthracene

Benzo[g,h,i]perylene
Pyrene

Retene
Fluoranthene

Chrysene

Perylene
different small scale heating appliances, a fireplace and a woodstove.
The authors found contributions to the PM2.5 mass ranging from 2.0 to
7.3 wt% and from 1.4 to 1.7 wt% in the experiments using the wood-
stove and the fireplace, respectively. In their study, the water-soluble
potassium comprised, on average, 33% and 21% of the analysed in-
organic ions in particles emitted from the fireplace and the woodstove,
respectively. In the present study, the contribution of potassium to the Fig. 6. PAH mass fractions in PM2.5 organic extracts from charcoal combustion.
total mass of water soluble ions was, on average, 34.0 ± 4.49%. Po-
tassium has been used as a tracer of biomass burning emissions. Alves multicyclic compounds. n‑Alkanes were in the range from C11 to C17
et al. [38] studied the composition of fine particles resulting from re- and from C23 to C27, maximising at C15, and accounted for 18.4 μg g−1
sidential wood combustion and observed that chloride and potassium PM2.5. As regards n‑alkenes, the highest mass fraction was observed for
were the major ionic compounds in the particulate matter samples. the C23 homologue (4.54 μg g−1 PM2.5), followed by C14 (2.05 μg g−1
PM2.5) and C16 (1.52 μg g−1 PM2.5). Squalene was also quantified, re-
3.3.2. Carbonaceous content presenting 0.61 μg g−1 PM2.5. This polyene was previously detected in
The carbonaceous matter represented 48.6 ± 4.05 wt% of the fireplace particle emissions from burning wood from six species of trees
PM2.5 mass emitted. The carbonaceous content was dominated by OC grown in the northeastern US [41]. Hopanes, triterpenoid hydrocarbons
(3.25 ± 0.384 g kg−1 of dry charcoal burned), which was found to formed through diagenetic and catagenetic processes over geological
correlate positively with the temperature above the fixed bed of fuel times during deep sediment burial, were not detected. Although present
(R2 = 0.832). A similar OC emission (4.8 g kg−1, of dry charcoal in coals with different degrees of geological maturity [42], the modern
burned) was reported by Andreae and Merlet [32]. The EC mass frac- and very immature carbonised wood of the present study may explain
tions in particles were substantially lower (0.342 ± 0.081 g kg−1 of the absence of these compounds from emissions. Steranes, which are
dry charcoal burned) and increased with increasing rates of fuel con- derived from the sterols of cell membranes of eukaryotes, mainly higher
version (R2 = 0.966). Vicente and Alves [39] reported OC and EC plants [43], were detected using the characteristic fragment at m/z 217.
particulate mass fractions from wood combustion in fireplaces in the As a whole, they accounted for 5.35 μg g−1 PM2.5.
range from 20 to 75 wt% and from 1.1 to 17 wt%, respectively. Vicente In total, the quantified PAHs constituted a PM2.5 mass fraction of
et al. [40] documented total carbonaceous contents ranging from 54.0 88.9 μg g−1 (Fig. 6). Retene represented 51% of this global mass frac-
to 73.0 wt% during hardwood and softwood combustion in a wood- tion. This aromatised derivative of resin acids has been found to be a
stove, under distinct operating conditions. good molecular marker of softwood combustion [44]. This over-
The OC/EC ratios have been used to describe the combustion con- whelming proportion of retene in the organic extracts of the present
ditions, with high combustion efficiency characterised by low OC/EC. study suggests that coal was manufactured, at least in part, from conifer
In the present study, an OC/EC ratio of 9.96 ± 3.03 was obtained. A wood. Following retene, the two most abundant polyaromatics were
positive correlation between the OC/EC ratio and temperature in the fluoranthene and pyrene, both 4-ring PAHs with MW 202. Fabiańska
combustion chamber (R2 = 0.836) and a negative correlation with the et al. [45] concluded that fly ashes from bituminous coal combustion in
fuel conversion rate (R2 = 0.985) were found. Vicente et al. [40] re- different types of boilers are rich in 4-ring PAHs (approximately 70%),
ported OC/EC ratios ranging from 1.1 to 6.1 and from 1.1 to 3.4 for whereas the source fuel (i.e., raw bituminous coal) is rich in 3-ring
softwood and hardwood combustion in a woodstove, respectively. For PAHs. Moreover, alkylphenanthrenes were found to be abundant
wood combustion in small scale combustion appliances, Calvo et al. compounds in these fly ashes. Methylphenanthrenes (m/z 192) and
[37] found OC/EC ratios ranging from 9.9 to 15.0 and from 3.2 to 10.1, dimethylphenanthrenes (m/z 206) were also detected in the organic
for fireplace and woodstove combustion, respectively. Even higher ra- extracts of charcoal combustion, representing mass fractions of 14.0
tios were reported by Alves et al. [38] for traditional residential wood and 102 μg g−1 PM2.5, respectively. Benzo[a]pyrene (BaP) is believed
burning appliances (14.4 ± 7.2 and 16.4 ± 7.6 for fireplace and to be the most toxic PAH, but other aromatics may additionally act as
woodstove, respectively). Thus, the OC/EC ratio is highly variable since carcinogens. Benzo[a]pyrene equivalent (BaPeq) concentrations were
it is closely related to combustion process parameters, fuels employed determined through multiplication of the individual PAH levels by their
and combustion technology used [39]. corresponding toxic equivalent factor [46,47]. A BaPeq mass fraction of
3.74 μg g−1 PM2.5 was obtained, accounting for 4.3% of total PAHs.
3.3.3. Organic speciation It is well known that PAH emissions greatly depend on the com-
The organic compounds identified in the solvent soluble fraction of bustion efficiency. It has been argued that emissions from coal-fired
PM2.5 emitted from charcoal combustion included, among others, ali- power plants are characterised by the absence of five or more ring
phatic compounds, PAHs, various types of acids, phenolic constituents, PAHs, while phenanthrene, fluorene, and fluoranthene dominate [48].
and anhydrosugars. Shen et al. [49] measured PAHs in PM2.5 emissions from residential
The aliphatic fraction comprised n‑alkanes, n‑alkenes and

301
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

cookstoves following the water boiling test protocol. A variety of fuel 0.332 mg g−1 PM2.5. Furthermore, several oxy-, hydroxyl- and diacids
and cookstove combinations were examined, including: (i) LPG, (ii) were distinguished. It has been documented that biomass burning is a
kerosene in a wick stove, (iii) wood in a forced-draft fan stove, and (iv) significant source of dicarboxylic acids and related compounds, such as
wood in a natural-draft rocket cookstove. The wood burning in the ketocarboxylic acids [51]. Among acids, diterpenoids especially dehy-
natural-draft stove had the highest PAH emissions followed by the droabietic and abietic acids, were the most abundant. These resin
wood combustion in the forced-draft stove and kerosene burning. The compounds were found to be major constituents of fine particles from
highest thermal efficiency and the lowest PAH emissions were found for wildfires in conifer-dominated forests (e.g., [28]), as well from the re-
LPG combustion. Compared with wood combustion, LPG burning also sidential combustion of softwood (e.g., [44]).
emitted a lower fraction of higher molecular weight PAH. The lowest In the present study, pinic acid was found to be one of the most
∑PAH25/PM2.5 fractions were observed for wood combustion in the important compounds in particulate matter emitted from charcoal
forced-draft stoves (1.13 ± 0.63 and 1.41 ± 0.88 μg mg−1 in high- combustion. It has been described as a secondary organic aerosol pro-
and low-moisture wood combustion during high-power phase and duct from the ozonolysis of α‑pinene (e.g., [52–55]). Its detection in the
2.81 ± 1.85 μg mg−1 in the low-power phase of dry wood combus- present study suggests that pinic acid forms in association with charcoal
tion), while the highest mass fractions were obtained for kerosene burning through pathways that are not clear and deserve further in-
burning in both the high- and low-power phases (10.1 ± 3.3 and vestigation. Possible mechanisms include the direct emission due to its
9.15 ± 5.44 μg mg−1, respectively). Shen et al. [50] characterised formation during the combustion of charcoal and/or emission of gas-
PAHs from the residential combustion of crop residues, woody material, eous precursors which are immediately oxidised to form the terpenic
coal, and biomass pellets in domestic stoves in rural China. The EFs of acid in the smoke plume. A few studies have documented that large
gaseous and particulate phase PAHs varied from several mg kg−1 for amounts of α- and β‑pinenes are emitted from combustion of foliar fuels
wood pellets, corn pellets, wood log, and anthracite briquettes to about and from prescribed burning in pine-dominated forests [56,57]. The
200 mg kg−1 for bituminous coal, showing a variation of almost two possibility that pinenes may be emitted from combustion of leaves and
orders of magnitude. Bituminous coals and brushwood yielded rela- trunks of softwood and related to the thermal decomposition of cellu-
tively large fractions of high molecular PAHs. The emission factor of lose in forest fires and from combustion of softwood logs in domestic
BaPeq for raw bituminous coal was as high as 52 mg kg−1, which was fireplaces and/or stoves has been argued by Cheng et al. [58]. These
1–2 orders of magnitude higher than the other fuels. The BaPeq EF researchers observed a consistent relationship between products from
obtained in the present study (27.6 μg kg−1) is of the same order of that the oxidation of pinenes and biomass burning tracers (e.g. levoglu-
reported by Shen et al. [50] for wood pellets (89 μg kg−1). cosan) in PM2.5 samples collected in wintertime in Toronto, Canada,
PAH isomer ratios are frequently used as diagnostic signatures for when the ozone and temperature values were minimal, thus, raising the
source apportionment by assuming that they differ significantly be- hypothesis of formation of terpenic acids during the wood combustion
tween sources and remain constant during transport to receptors. Shen processes.
et al. [50] concluded that, for source diagnosis, high molecular weight In the present study, anhydrosugars constituted the dominant or-
isomers are more informative than low molecular weight ones and ganic class in PM2.5. Levoglucosan (L) and its isomers (mannosan and
multiple ratios could be used together whenever possible. The ratios galactosan, M and G) are formed from the thermal breakdown of woody
obtained in this study (Table 4) are, in general, consistent with coal and biomass constituents at relatively low temperatures. In addition to
biomass burning emissions, although they cannot differentiate between dissimilar cellulosic contents of the different biofuels, the emission of
these two sources. However, the BaP/(BaP + BghiP) ratio overlaps with saccharidic compounds may strongly depend on the combustion
that reported for traffic sources. Thus, PAH isomer ratios should be equipment and operating conditions [39]. Levoglucosan emissions of
interpreted with care, avoiding their application as a standalone 90 ± 31, 123 ± 64, 56 ± 42 and 1.8 ± 1.2 mg g−1 PM have been
methodology for diagnostic signatures in source apportionment studies. reported for the combustion of biofuels in a fireplace (PM2.5), tradi-
Some n‑alkanols, representing 2.65 mg g−1 PM2.5, were detected in tional woodstove (PM2.5), ecolabelled woodstove (PM10) and pellet
the organic extracts. Among them, octadecanol deserves attention as stove (PM2.5), respectively ([59], and references therein). Sheesley
the most abundant (Table 5). A series of low molecular weight n‑alk- et al. [60] reported values of 27.9 ± 5.3, 18.3 ± 3.5, 19.1 ± 3.6,
anoic acids from C8 to C26 with Cmax at C16, C22 and C24, was also ob- 98.4 ± 18.7 and 44.5 ± 8.5 mg g−1 PM2.5 for levoglucosan emissions
served in the PM2.5 organic extracts, accounting for 1.35 mg g−1. Un- from the combustion of coconut leaves, rice straw, cow dung, biomass
saturated fatty acids (C16:1, C18:1 and C18:2) represented briquettes and jackfruit branches, respectively. Although the three

Table 4
Diagnostic ratios between PAHs and comparison with values from the literature.

Ratio This study Reference values Interpretation References (and citations therein)

Ant/(Ant + Phen) 0.12 < 0.1 Petrogenic sources [50]


> 0.1 Pyrogenic sources
Flu/(Flu + Pyr) 0.47 < 0.4 Petrogenic sources [47]
0.4–0.5 Fossil fuel combustion
> 0.5 Grass, wood and coal combustion
BaA/(BaA + Chry) 0.47 < 0.2 Petrogenic sources [71]
0.2–0.35 Mixture of petrogenic and combustion sources
> 0.5 Combustion sources
IcdP/(IcdP + BghiP) 0.50 0.2 Petroleum [47,71]
0.2–0.5 Liquid fossil fuel combustion
> 0.5 Grass, wood and coal combustion
BbF/(BbF + BkF) 0.53 0.53 Residential wood pellet stove [46]
0.55 Residential straw pellet stove
0.60–0.89 Residential coal stove
BaP/(BaP + BghiP) 0.57 > 0.38 Traffic sources [50]
< 0.38 Non-traffic sources

Ant – anthracene, Phen – phenanthrene, Flu – fluoranthene, Pyr – pyrene, BaA - benzo[a]anthracene, Chry – chrysene, BaP - benzo[a]pyrene, BghiP - benzo[g,h,i]perylene.

302
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

Table 5 Table 5 (continued)


Oxygen containing organic compounds in PM2.5 emitted from charcoal combustion in a
brick barbecue. PM2.5 mass Chemical
fractions formula
PM2.5 mass Chemical (mg g−1)
fractions formula
(mg g−1) 1,5‑Pentanedioic (glutaric) 0.404 C5H8O4
Hexanedioic (adipic) 0.030 C6H10O4
Saccharides Heptanedioic (pimelic) 0.063 C7H12O4
Levoglucosan 107 C6H10O5 Octanedioic (suberic) 0.116 C8H14O4
Mannosan 34.2 C6H10O5 Nonanedioic (azelaic) 0.051 C9H16O4
Galactosan 8.59 C6H10O5 Decanedioic (sebacic) 0.004 C10H18O4
Maltose 7.60 C12H22O11 Hexadecanedioic (thapsic) 0.019 C16H30O4
Other saccharides 56.7 – Pinonic 0.058 C9H14O4
Pinic 3.39 C9H14O4
Phenolic compounds
Phthalic 0.468 C8H6O4
Sinapic acid 0.003 C11H12O5
Cinnamic 0.026 C9H8O2
Sinapyl alcohol 1.65 C11H14O4
Benzoic 0.015 C7H6O2
Coniferyl alcohol 12.1 C10H12O3
Homovanillic acid 1.70 C9H10O4 n‑Alkanols
3‑Vanillylpropanol 2.52 C11H16O3 1‑Pentadecanol 0.002 C15H32O
Homovanillyl alcohol 0.216 C9H12O3 1‑Docosanol 0.291 C22H46O
Vanillic acid 1.13 C8H8O4 1‑Tricosanol 0.027 C23H48O
Vanillylpropionic acid 0.425 C10H12O4 1‑Hexacosanol 0.105 C26H54O
2,4‑di‑tert‑Butylphenol 0.106 C14H22O 1‑Heptacosanol 0.051 C27H56O
2,6‑Dimethoxyphenol (syringol) 0.025 C8H10O3 1‑Octacosanol 2.12 C28H58O
Syringic acid 0.283 C9H10O5 1‑Tricontanol 0.05 C30H62O
Acetosyringone 0.169 C10H12O4
Other hidroxyl compounds and lactones
3,4‑Dihydroxy‑cinnamic (caffeic) acid 0.004 C9H8O4
Glycerol 0.363 C3H8O3
Hydroxycinnamic (coumaric) acid 0.462 C9H8O3
1‑Monolauroyl‑rac‑glycerol 0.011 C15H30O4
3‑Methoxy‑4‑hydroxycinnamic (ferulic) acid 0.033 C10H10O4
Citronellol 0.004 C10H20O
Hydroxybenzene derivative 0.032 –
2,3‑Pinanediol 0.010 C10H18O2
4‑Hydroxybenzoic acid 0.237 C7H6O3
2‑Methylcyclohexanol 0.001 C7H14O
3,4,5‑Trihydroxybenzoic (gallic) 0.020 C7H6O5
5‑Cholestan‑3‑ol 0.015 C27H48O
Hydroquinone 0.003 C6H6O2
β‑Sitosterol 1.60 C29H50O
Catechol 0.543 C6H6O2
L‑Threonate‑1,4‑lactone 0.102 C4H6O4
4‑Hydroxyacetophenone 0.009 C8H8O2
1,2,3,4,5‑Pentahydroxycaproic acid δ‑lactone 0.066 C6H10O6
Methylcatechol 0.369 C7H8O2
(δ‑Gluconolactone)
Resorcinol 0.517 C6H6O2
3‑Deoxy‑D‑ribo‑hexono‑1,4‑lactone 2.85 C6H10O5
Pyrogallol 2.00 C6H6O3
Tocopherol 0.692 C29H50O2
4‑Phenylphenol 0.103 C12H10O
4‑Octyphenol 0.367 C14H22O isomers are simultaneously emitted by biomass burning sources, their
share varies greatly with the type of fuel, such as softwood versus
n‑Alkenoic and n‑alkanoic acids
9‑Hexadecenoic (palmitoleic) 0.026 C16H30O2 hardwood. Fabbri et al. [61] demonstrated that burning of lignites is an
9‑12‑Octadecadienoic (linoleic) 0.147 C18H32O2 additional source of levoglucosan to the atmosphere in regions where
9‑Octadecenoic (oleic) 0.159 C18H34O2 brown coal is used as a domestic fuel. Galactosan was not emitted in
Octanoic (caprylic) 0.010 C8H16O2
lignite smoke, while mannosan was present at relatively low con-
Nonanoic (pelargonic) 0.010 C9H18O2
Decanoic (capric) 0.001 C10H20O2
centrations. Thus, the same researchers proposed L/M and L/(M + G)
Undecanoic acid 0.028 C11H22O2 as discriminator ratios between products from combustion of lignites
Dodecanoic (lauric) 0.046 C12H24O2 and biomass. Values from 31 to 92 were obtained in emissions from
Tetradecanoic (myristic) 0.001 C14H28O2 burning lignite [61]. L/M and L/(M + G) ratios of 3.1 and 2.5, re-
Pentadecanoic 0.020 C15H30O2
spectively, were found in PM2.5 of the present study, which compare
Hexadecanoic (palmitic) 0.273 C16H32O2
Heptadecanoic 0.007 C17H34O2 with ratios of 2.5 and 2.0 reported for charred pine wood [62] and the
Octadecanoic 0.038 C18H36O2 range of values compiled from the bibliography by Fabbri et al. [61] for
Nonadecanoic 0.003 C19H38O2 softwood burning. However, the wide spectrum of values found in
Eicosanoic 0.022 C20H40O2
emissions from the combustion of grasses for L/M (2.0 to 33. 3) and L/
Docosanoic 0.377 C22H44O2
Tricosanoic 0.037 C23H46O2
(M + G) (1.7 to 9.5) [63], as well as the low ratios (3.3–8.4 and
Tetracosanoic 0.303 C24H48O2 1.5–2.1, respectively) documented for hardwood combustion [64], can
Hexacosanoic 0.171 C26H52O2 lead to question the discriminatory potential of these ratios, due to
Resin acids overlap between biomass sources.
Abietic acid 4.80 C20H30O2 The most abundant components of biomass include cellulose,
Dehydroabietic acid 10.8 C20H28O2 hemicelluloses and lignin. The latter is a biopolymer mainly composed
Isopimaric acid 3.06 C20H30O2
of alkylphenols. During combustion, lignin yields a wide range of
Podocarpic acid 0.010 C17H22O3
products, of which the most characteristic ones are methoxy phenols
Other acids
[39]. Gymnosperms, as conifers, have a lignin that consists almost en-
Glycolic 5.21 C2H4O3
Dioxobutanoic 0.242 C4H4O4
tirely of coniferyl alcohol. Dicotyledonic lignin is a mixture of coniferyl
2‑Hydroxybutyric 0.129 C4H8O3 alcohol and sinapyl alcohol, whereas monocotyledonic lignin is a
3‑Hydroxybutyric 0.010 C4H8O3 mixture of three monolignols (coniferyl, sinapyl and coumaryl alcohols)
3‑Hydroxypropanoic (hydracrylic) 0.180 C3H6O3 [65]. Some monocotyledons have mostly coniferyl alcohol (as many
4‑Oxopentanoico (levulinic) 0.061 C5H8O3
grasses), while others have mainly sinapyl alcohols. These three lignols
2‑Furoic 0.399 C5H4O3
Pyrotartaric 0.227 C5H8O4 (coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol) are in-
Propanedioic (malonic) 1.044 C3H4O4 corporated into lignin in the form of the phenylpropanoids p‑hydrox-
yphenyl, guaiacyl, and syringyl, respectively [65]. Coniferyl alcohol

303
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

was the most abundant phenolic compound in PM2.5 from charcoal anhydrosugars were much lower than the values reported for lignite
combustion in this study (Table 5). This suggests that charcoal was combustion, but overlap those from other biomass emission sources.
made of a mixture of mixture of hardwood and softwood. The most Thus, it is not possible the discrimination between emissions of these
common biomass species used for charcoal production in Portugal are compounds from charcoal and other biomass combustion sources.
angiosperms (hardwood), such as eucalypt, olive tree and cork oak. Likewise, although some isomeric ratios between PAHs are consistent
Although to a lesser extent, pine (softwood) is also used. Apart from the with biomass and coal combustion, specific ratios do not allow to dis-
charcoal itself, the wood added at the beginning of the combustion as tinguish between petrogenic (oil-based) and pyrogenic (combustion-
kindling may have had some contribution to the measured concentra- based) sources. Thus, to diagnose sources, the application of multiple
tions. Coumaryl (e.g. coumaric acid), sinapyl (e.g. sinapyl alcohol) and ratios, together with other assignment tools, is recommended.
guaiacyl derivatives (e.g. vanillic acid, homovanillic acid and vanillyl-
propanol) were other dominant phenolic compounds in the PM2.5 Acknowledgments
emitted during the charcoal combustion experiments in this work
(Table 5). The presence of a significant amount of pyrogallol is also Thanks are given to the POHP/FSE programme and to the
noteworthy. High PM2.5 mass fractions of this benzenetriol were pre- Portuguese Foundation of Science and Technology (FCT) for funding
viously reported by Gonçalves et al. [44] for the combustion of several the scholarships SFRH/BPD/88988/2012, SFRH/BD/117993/2016 and
Portuguese wood species in a stove and a fireplace. SFRH/BPD/123176/2016. An acknowledgment is also due for the fi-
Only two sterols, cholestanol and β‑sitosterol, were identified in the nancial support to CESAM (UID/AMB/50017), to FCT/MEC through
organic extracts from the charcoal combustion in this study. national funds, and the co-funding by the FEDER, within the PT2020
Cholestanol, the less abundant, has been described as a faecal stanol Partnership Agreement and Compete 2020.
and pointed out as a representative molecular marker of the products of
cow's manure burning [66]. Sheesley et al. [60] found this C27 sterol in References
PM2.5 emissions from the combustion of cow dung
(1.63 ± 0.53 mg g−1), but did not detect it in particulate matter [1] FAO, Forest Products Statistics. 2014 Global Forest Products Facts and Figures,
samples from wood or foliar fuels combustion. However, cholestanol (2015), http://dx.doi.org/10.1038/1821553c0.
[2] M.J. Antal, M. Grønli, The art, science, and technology of charcoal production, Ind.
has also been reported as an extractive component of some woods Eng. Chem. Res. 42 (2003) 1619–1640, http://dx.doi.org/10.1021/ie0207919.
[67,68]. Moreover, its biosynthesis from both cholesterol (C27 sterol) [3] E. Johnson, Charcoal versus LPG grilling: a carbon-footprint comparison, Environ.
and campestanol (C28 sterol) in higher plants has been documented Impact Assess. Rev. 29 (2009) 370–378, http://dx.doi.org/10.1016/j.eiar.2009.02.
004.
[69]. Thus, the possibility of having multiple origins is a likely scenario. [4] S. Bonjour, H. Adair-Rohani, J. Wolf, N.G. Bruce, S. Mehta, A. Prüss-Ustün,
β‑Sitosterol, the most common steroid in wood and higher plants, was M. Lahiff, E.A. Rehfuess, V. Mishra, K.R. Smith, Solid fuel use for household
one of the dominant polar organic compounds in PM2.5 from charcoal cooking: country and regional estimates for 1980–2010, Environ. Health Perspect.
121 (2013) 784–790, http://dx.doi.org/10.1289/ehp.1205987.
combustion. It also occurred in emitted fly ashes from bituminous coal [5] J.C. Adam, Improved and more environmentally friendly charcoal production
and bituminous coal and biomass combustion in various boilers [45]. system using a low-cost retort-kiln (eco-charcoal), Renew. Energy 34 (2009)
β‑Sitosterol was described as the most abundant phytosterol in parti- 1923–1925, http://dx.doi.org/10.1016/j.renene.2008.12.009.
[6] E.N. Chidumayo, D.J. Gumbo, The environmental impacts of charcoal production in
culate matter emitted from fireplace and stove combustion of several
tropical ecosystems of the world: a synthesis, Energy Sustain. Dev. 17 (2013) 86–94,
wood species, among which the highest concentrations were obtained http://dx.doi.org/10.1016/j.esd.2012.07.004.
for pine [44]. [7] M. Sparrevik, G. Cornelissen, M. Sparrevik, C. Adam, V. Martinsen, G. Cornelissen,
Several carbohydrate-based lactones were present in the organic G. Cornelissen, Emissions of gases and particles from charcoal/biochar production
in rural areas using medium-sized traditional and improved “retort” kilns, Biomass
extracts. However, only a few could be quantified due to the lack of Bioenerg. 72 (2015) 65–73, http://dx.doi.org/10.1016/j.biombioe.2014.11.016.
authentic standards and the similarity of mass spectral fragmentation [8] S.K. Akagi, R.J. Yokelson, C. Wiedinmyer, M.J. Alvarado, J.S. Reid, T. Karl,
patterns. As far as we know, this is the first time that these compounds J.D. Crounse, P.O. Wennberg, Emission factors for open and domestic biomass
burning for use in atmospheric models, Atmos. Chem. Phys. 11 (2011) 4039–4072,
are identified in particulate matter emitted from biofuel combustion. http://dx.doi.org/10.5194/acp-11-4039-2011.
Charring of cellulose appears to be responsible for the creation of car- [9] D.M. Kammen, D.J. Lew, Review of technologies for the production and use of
boxylic functional groups and lactones [70]. Further research is needed charcoal, Renew. Appropr. Energy Lab. Rep. 2005, pp. 1–19.
[10] D.M. Pennise, K.R. Smith, J.P. Kithinji, M.E. Rezende, T.J. Raad, J. Zhang, C. Fan,
to broaden the range of compounds and to clarify their formation me- Emissions of greenhouse gases and other airborne pollutants from charcoal making
chanisms. in Kenya and Brazil, J. Geophys. Res. 106 (2001) 24143, , http://dx.doi.org/10.
1029/2000JD000041.
[11] C.Y. Kuo, H.S. Lee, J.H. Lai, Emission of polycyclic aromatic hydrocarbons and lead
4. Conclusions
during Chinese mid-autumn festival, Sci. Total Environ. 366 (2006) 233–241,
http://dx.doi.org/10.1016/j.scitotenv.2005.08.006.
Emissions of carbon oxides, NOx, acid gases, NH3 and gas-phase [12] C.A. Alves, M. Duarte, T. Nunes, R. Moreira, S. Rocha, Carbonaceous particles
emitted from cooking activities in Portugal, Glob. NEST J. 16 (2014) 412–420.
TOC from the combustion of charcoal in a typical brick barbecue grill
[13] C.A. Alves, M. Evtyugina, M. Cerqueira, T. Nunes, M. Duarte, E. Vicente, Volatile
were obtained and compared to those of other fuels and appliances. organic compounds emitted by the stacks of restaurants, Air Qual. Atmos. Health 8
Methane represented the highest fraction (73.2%) of TOC emissions, (2015) 401–412, http://dx.doi.org/10.1007/s11869-014-0310-7.
followed by ethane (14.9%) and hexane (5.9%). Particle emissions were [14] S.C. Lee, W.M. Li, L. Yin Chan, Indoor air quality at restaurants with different styles
of cooking in metropolitan Hong Kong, Sci. Total Environ. 279 (2001) 181–193,
of the same order of magnitude as those from traditional residential http://dx.doi.org/10.1016/S0048-9697(01)00765-3.
wood burning appliances. The mass of PM2.5 in the combustion flue gas [15] J.-B. Lee, K.-H. Kim, H.-J. Kim, S.-J. Cho, K. Jung, S.-D. Kim, Emission rate of
was composed mainly of organic matter. Water soluble ions accounted particulate matter and its removal efficiency by precipitators in under-fired char-
broiling restaurants, Sci. World J. 11 (2011) 1077–1088, http://dx.doi.org/10.
for around 17% of the PM2.5 mass. Potassium, sodium and chloride 1100/tsw.2011.103.
were the dominant water soluble inorganic ions in PM2.5. Although [16] X. Hou, G. Zhuang, Y. Lin, J. Li, Y. Jiang, J.S. Fu, Emission of fine organic aerosol
quantitative information for organic compounds has been previously from traditional charcoal broiling in China, J. Atmos. Chem. 61 (2008) 119–131,
http://dx.doi.org/10.1007/s10874-009-9128-3.
reported for wood combustion in residential systems and other coal [17] S. Taner, B. Pekey, H. Pekey, Fine particulate matter in the indoor air of barbeque
types, as far as we know, this is the first time that a comprehensive restaurants: elemental compositions, sources and health risks, Sci. Total Environ.
speciation of PM2.5 emitted from charcoal combustion is obtained. The 454-455 (2013) 79–87, http://dx.doi.org/10.1016/j.scitotenv.2013.03.018.
[18] H.L. Huang, W.M.G. Lee, F.S. Wu, Emissions of air pollutants from indoor charcoal
new database may be useful to support source apportionment studies barbecue, J. Hazard. Mater. 302 (2016) 198–207, http://dx.doi.org/10.1016/j.
and to improve emission inventories. The overwhelming proportion of jhazmat.2015.09.048.
retene, resin acids, coniferyl and vanillyl compounds is consistent with [19] E. Kabir, K.H. Kim, J.W. Ahn, O.F. Hong, J.R. Sohn, Barbecue charcoal combustion
as a potential source of aromatic volatile organic compounds and carbonyls, J.
emissions from charred coniferous wood. The ratios between

304
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

Hazard. Mater. 174 (2010) 492–499, http://dx.doi.org/10.1016/j.jhazmat.2009. C. Pio, Organic compounds in PM2.5 emitted from fireplace and woodstove com-
09.079. bustion of typical Portuguese wood species, Atmos. Environ. 45 (2011) 4533–4545,
[20] M. Olsson, G. Petersson, Benzene emitted from glowing charcoal, Sci. Total Environ. http://dx.doi.org/10.1016/j.atmosenv.2011.05.071.
303 (2003) 215–220, http://dx.doi.org/10.1016/S0048-9697(02)00403-5. [45] M. Fabiańska, B. Kozielska, J. Konieczyński, Differences in the occurrence of
[21] M.M. Rahman, K.H. Kim, Release of offensive odorants from the combustion of polycyclic aromatic hydrocarbons and geochemical markers in the dust emitted
barbecue charcoals, J. Hazard. Mater. 215-216 (2012) 233–242, http://dx.doi.org/ from various coal-fired boilers, Energ. Fuel. 31 (2017) 2585–2595, http://dx.doi.
10.1016/j.jhazmat.2012.02.055. org/10.1021/acs.energyfuels.6b03030.
[22] E. Kabir, K.H. Kim, H.O. Yoon, Trace metal contents in barbeque (BBQ) charcoal [46] X. Yang, C. Geng, X. Sun, W. Yang, X. Wang, J. Chen, Characteristics of particulate-
products, J. Hazard. Mater. 185 (2011) 1418–1424, http://dx.doi.org/10.1016/j. bound polycyclic aromatic hydrocarbons emitted from industrial grade biomass
jhazmat.2010.10.064. boilers, J. Environ. Sci. (China) 40 (2016) 28–34, http://dx.doi.org/10.1016/j.jes.
[23] J. Susaya, K.H. Kim, J.W. Ahn, M.C. Jung, C.H. Kang, BBQ charcoal combustion as 2015.09.010.
an important source of trace metal exposure to humans, J. Hazard. Mater. 176 [47] S. Bootdee, S. Chantara, T. Prapamontol, Determination of PM2.5 and polycyclic
(2010) 932–937, http://dx.doi.org/10.1016/j.jhazmat.2009.11.129. aromatic hydrocarbons from incense burning emission at shrine for health risk
[24] A. Dyremark, R. Westerholm, E. Overvik, J.-A. Gustavsson, Polycyclic aromatic assessment, Atmos. Pollut. Res. 7 (2016) 680–689, http://dx.doi.org/10.1016/j.
hydrocarbon (PAH) emissions from charcoal grilling, Atmos. Environ. 29 (1995) apr.2016.03.002.
1553–1558, http://dx.doi.org/10.1016/1352-2310(94)00357-Q. [48] C.C. Revuelta, E.d. la F. Santiago, J.A.R. Vázquez, Characterization of polycyclic
[25] N.T.K. Oanh, L.B. Reutergårdh, N.T. Dung, Emission of polycyclic aromatic hy- aromatic hydrocarbons in emissions from coal-fired power plants: the influence of
drocarbons and particulate matter from domestic combustion of selected fuels, operation parameters, Environ. Technol. 20 (1999) 61–68.
Environ. Sci. Technol. 33 (1999) 2703–2709, http://dx.doi.org/10.1021/ [49] G. Shen, W. Preston, S.M. Ebersviller, C. Williams, J.W. Faircloth, J.J. Jetter,
es980853f. M.D. Hays, Polycyclic aromatic hydrocarbons in fine particulate matter emitted
[26] N. MacCarty, D. Still, D. Ogle, Fuel use and emissions performance of fifty cooking from burning kerosene, liquid petroleum gas, and wood fuels in household cook-
stoves in the laboratory and related benchmarks of performance, Energy Sustain. stoves, Energ. Fuel. 31 (2017) 3081–3090, http://dx.doi.org/10.1021/acs.
Dev. 14 (2010) 161–171, http://dx.doi.org/10.1016/J.ESD.2010.06.002. energyfuels.6b02641.
[27] A.I. Calvo, L.A.C. Tarelho, C.A. Alves, M. Duarte, T. Nunes, Characterization of [50] G. Shen, S. Tao, Y. Chen, Y. Zhang, S. Wei, M. Xue, B. Wang, R. Wang, Y. Lu, W. Li,
operating conditions of two residential wood combustion appliances, Fuel Process. H. Shen, Y. Huang, H. Chen, Emission characteristics for polycyclic aromatic hy-
Technol. 126 (2014) 222–232, http://dx.doi.org/10.1016/j.fuproc.2014.05.001. drocarbons from solid fuels burned in domestic stoves in rural China, Environ. Sci.
[28] C.A. Alves, A. Vicente, C. Monteiro, C. Gonçalves, M. Evtyugina, C. Pio, Emission of Technol. 47 (2013) 14485–14494, http://dx.doi.org/10.1021/es403110b.
trace gases and organic components in smoke particles from a wildfire in a mixed- [51] S. Kundu, K. Kawamura, T.W. Andreae, A. Hoffer, M.O. Andreae, Molecular dis-
evergreen forest in Portugal, Sci. Total Environ. 409 (2011) 1466–1475, http://dx. tributions of dicarboxylic acids, ketocarboxylic acids and α-dicarbonyls in biomass
doi.org/10.1016/j.scitotenv.2010.12.025. burning aerosols: implications for photochemical production and degradation in
[29] J.J. Jetter, P. Kariher, Solid-fuel household cook stoves: Characterization of per- smoke layers, Atmos. Chem. Phys. Discuss. 9 (2010) 19783–19815, http://dx.doi.
formance and emissions, Biomass Bioenerg. 33 (2008) 294–305, http://dx.doi.org/ org/10.5194/acpd-9-19783-2009.
10.1016/j.biombioe.2008.05.014. [52] Y. Ma, T.R. Willcox, A.T. Russell, G. Marston, Pinic and pinonic acid formation in
[30] J. Jetter, Y. Zhao, K.R. Smith, B. Khan, T. Yelverton, P. Decarlo, M.D. Hays, the reaction of ozone with α-pinene, Chem. Commun. (2007) 1328–1330, http://
Pollutant emissions and energy efficiency under controlled conditions for house- dx.doi.org/10.1039/B617130C.
hold biomass cookstoves and implications for metrics useful in setting international [53] T.S. Christoffersen, J. Hjorth, O. Horie, N.R. Jensen, D. Kotzias, L.L. Molander,
test standards, Environ. Sci. Technol. 46 (2012) 10827–10834, http://dx.doi.org/ P. Neeb, L. Ruppert, R. Winterhalter, A. Virkkula, K. Wirtz, B.R. Larsen, cis-Pinic
10.1021/es301693f. acid, a possible precursor for organic aerosol formation from ozonolysis of α-
[31] S.C. Bhattacharya, D.O. Albina, P. Abdul Salam, Emission factors of wood and pinene, Atmos. Environ. 32 (1998) 1657–1661, http://dx.doi.org/10.1016/S1352-
charcoal-fired cookstoves, Biomass Bioenerg. 23 (2002) 453–469, http://dx.doi. 2310(97)00448-2.
org/10.1016/S0961-9534(02)00072-7. [54] J. Warnke, R. Bandur, T. Hoffmann, Capillary-HPLC-ESI-MS/MS method for the
[32] M.O. Andreae, P. Merlet, Emission of trace gases and aerosols from biomass determination of acidic products from the oxidation of monoterpenes in atmo-
burning, Glob. Biogeochem. Cy. 15 (2001) 955–966. spheric aerosol samples, Anal. Bioanal. Chem. 385 (2006) 34–45, http://dx.doi.
[33] T. Brunner, I. Obernberger, R. Scharler, Primary measures for low-emission re- org/10.1007/s00216-006-0340-6.
sidential wood combustion – comparison of old with optimised modern systems, in: [55] J. Yu, R.J. Griffin, D.R. Cocker, R.C. Flagan, J.H. Seinfeld, P. Blanchard,
ETA-Renewable Energies (Ed.), 17th Eur. Biomass Conf. Exhib., Hamburg, 2009. Observation of gaseous and particulate products of monoterpene oxidation in forest
[34] R. Koppmann, K. von Czapiewski, J.S. Reid, A review of biomass burning emissions, atmospheres, Geophys. Res. Lett. 26 (1999) 1145–1148, http://dx.doi.org/10.
part I: gaseous emissions of carbon monoxide, methane, volatile organic com- 1029/1999GL900169.
pounds, and nitrogen containing compounds, Atmos. Chem. Phys. Discuss. 5 (2005) [56] M.D. Hays, C.D. Geron, K.J. Linna, N.D. Smith, J.J. Schauer, Speciation of gas-phase
10455–10516, http://dx.doi.org/10.5194/acpd-5-10455-2005. and fine particle emissions from burning of foliar fuels, Environ. Sci. Technol. 36
[35] I.T. Bertschi, R.J. Yokelson, D.E. Ward, T.J. Christian, W.M. Hao, Trace gas emis- (2002) 2281–2295, http://dx.doi.org/10.1021/ES0111683.
sions from the production and use of domestic biofuels in Zambia measured by [57] S. Lee, K. Baumann, J.J. Schauer, R.J. Sheesley, L.P. Naeher, S. Meinardi,
open-path Fourier transform infrared spectroscopy, J. Geophys. Res. Atmos. 108 D.R. Blake, E.S. Edgerton, A.G. Russell, M. Clements, Gaseous and particulate
(2003) 8469, http://dx.doi.org/10.1029/2002JD002158. emissions from prescribed burning in Georgia, Environ. Sci. Technol. 39 (2005)
[36] C. Gonçalves, C. Alves, M. Evtyugina, F. Mirante, C. Pio, A. Caseiro, C. Schmidl, 9049–9056, http://dx.doi.org/10.1021/ES051583L.
H. Bauer, F. Carvalho, Characterisation of PM10 emissions from woodstove com- [58] Y. Cheng, J.R. Brook, S.M. Li, A. Leithead, Seasonal variation in the biogenic sec-
bustion of common woods grown in Portugal, Atmos. Environ. 44 (2010) ondary organic aerosol tracer cis-pinonic acid: enhancement due to emissions from
4474–4480, http://dx.doi.org/10.1016/j.atmosenv.2010.07.026. regional and local biomass burning, Atmos. Environ. 45 (2011) 7105–7112, http://
[37] A.I. Calvo, V. Martins, T. Nunes, M. Duarte, R. Hillamo, K. Teinilä, V. Pont, dx.doi.org/10.1016/j.atmosenv.2011.09.036.
A. Castro, R. Fraile, L. Tarelho, C. Alves, Residential wood combustion in two do- [59] C.A. Alves, E.D. Vicente, S. Rocha, A.M. Vicente, Organic tracers in aerosols from
mestic devices: relationship of different parameters throughout the combustion the residential combustion of pellets and agro-fuels, Air Qual. Atmos. Health 10
cycle, Atmos. Environ. 116 (2015) 72–82, http://dx.doi.org/10.1016/j.atmosenv. (2017) 37–45, http://dx.doi.org/10.1007/s11869-016-0406-3.
2015.06.012. [60] R.J. Sheesley, J.J. Schauer, Z. Chowdhury, G.R. Cass, B.R.T. Simoneit,
[38] C. Alves, C. Gonçalves, A.P. Fernandes, L. Tarelho, C. Pio, Fireplace and woodstove Characterization of organic aerosols emitted from the combustion of biomass in-
fine particle emissions from combustion of western Mediterranean wood types, digenous to South Asia, J. Geophys. Res. Atmos. 108 (2003), http://dx.doi.org/10.
Atmos. Res. 101 (2011) 692–700. 1029/2002JD002981 4285(D9).
[39] E.D. Vicente, C.A. Alves, An overview of particulate emissions from residential [61] D. Fabbri, C. Torri, B.R.T. Simoneit, L. Marynowski, A.I. Rushdi, M.J. Fabiańska,
biomass combustion, Atmos. Res. 199 (2017) 159–185, http://dx.doi.org/10.1016/ Levoglucosan and other cellulose and lignin markers in emissions from burning of
j.atmosres.2017.08.027. Miocene lignites, Atmos. Environ. 43 (2009) 2286–2295, http://dx.doi.org/10.
[40] E.D. Vicente, M.A. Duarte, A.I. Calvo, T.F. Nunes, L.A.C. Tarelho, D. Custódio, 1016/j.atmosenv.2009.01.030.
C. Colombi, V. Gianelle, A. Sanchez de la Campa, C.A. Alves, Influence of operating [62] A. Otto, R. Gondokusumo, M.J. Simpson, Characterization and quantification of
conditions on chemical composition of particulate matter emissions from residential biomarkers from biomass burning at a recent wildfire site in Northern Alberta,
combustion, Atmos. Res. 166 (2015) 92–100, http://dx.doi.org/10.1016/j. Canada, Appl. Geochem. 21 (2006) 166–183, http://dx.doi.org/10.1016/J.
atmosres.2015.06.016. APGEOCHEM.2005.09.007.
[41] P.M. Fine, G.R. Cass, B.R.T. Simoneit, Chemical characterization of fine particle [63] D.R. Oros, M.R. bin Abas, N.Y.M.J. Omar, N.A. Rahman, B.R.T. Simoneit,
emissions from fireplace combustion of woods grown in the northeastern United Identification and emission factors of molecular tracers in organic aerosols from
States, Environ. Sci. Technol. 35 (2001) 2665–2675, http://dx.doi.org/10.1021/ biomass burning: part 3. Grasses, Appl. Geochem. 21 (2006) 919–940, http://dx.
es001466k. doi.org/10.1016/J.APGEOCHEM.2006.01.008.
[42] F. Han, J. Cao, L. Peng, H. Bai, D. Hu, L. Mu, X. Liu, Characteristics of hopanoid [64] D.R. Oros, B.R. Simoneit, Identification and emission factors of molecular tracers in
hydrocarbons in ambient PM10 and motor vehicle emissions and coal ash in organic aerosols from biomass burning part 2. Deciduous trees, Appl. Geochem. 16
Taiyuan, China, Environ. Geochem. Health 37 (2015) 813–829, http://dx.doi.org/ (2001) 1545–1565, http://dx.doi.org/10.1016/S0883-2927(01)00022-1.
10.1007/s10653-015-9763-3. [65] A. Dermibas, Fuels from biomass, Biohydrogen. Futur. Engine Fuel Demands,
[43] A.S. Mackenzie, S.C. Brassell, G. Eglinton, J.R. Maxwell, Chemical fossils: the Springer-Verlag, London, 2009, pp. 43–59, , http://dx.doi.org/10.1007/978-1-
geological fate of steroids, Science 217 (1982) 491–504 http://science.sciencemag. 84882-511-6.
org/content/217/4559/491.abstract. [66] K.Y. Kondratyev, L.S. Ivlev, V.F. Krapivin, C.A. Varostos, Atmospheric Aerosol
[44] C. Gonçalves, C. Alves, A.P. Fernandes, C. Monteiro, L. Tarelho, M. Evtyugina, Properties: Formation, Processes and Impacts, Springer-Verlag, Germany, in

305
E.D. Vicente et al. Fuel Processing Technology 176 (2018) 296–306

association with Praxis Publishing Ltd., Chichester, UK, 2006. 1016/S0031-9422(02)00113-9.


[67] B. Jansen, K.G.J. Nierop, F.H. Tonneijck, F.W.M. Van Der Wielen, J.M. Verstraten, [70] D.W. Rutherford, R.L. Wershaw, J.B. Reeves III, Development of Acid Functional
Can isoprenoids in leaves and roots of plants serve as biomarkers for past vegetation Groups and Lactones During the Thermal Degradation of Wood and Wood
changes? A case study from the Ecuadorian Andes, Plant Soil 291 (2007) 181–198, Components, (2008).
http://dx.doi.org/10.1007/s11104-006-9185-1. [71] H. An, G. Zhang, C. Liu, H. Guo, W. Yin, X. Xia, Characterization of PM2.5-bound
[68] J. Rowe, A. Conner, Extractives in Esastern hardwoods, General Technical Report polycyclic aromatic hydrocarbons and its deposition in Populus tomentosa leaves in
FPL 18, Forest Products Laboratory, Madison, Wisconsin, 1979. Beijing, Environ. Sci. Pollut. Res. 24 (2017) 8504–8515, http://dx.doi.org/10.
[69] N. Nakajima, S. Fujioka, T. Tanaka, S. Takatsuto, S. Yoshida, Biosynthesis of cho- 1007/s11356-017-8516-5.
lestanol in higher plants, Phytochemistry 60 (2002) 275–279, http://dx.doi.org/10.

306

You might also like