Sunteți pe pagina 1din 13

Atmospheric Environment 120 (2015) 15e27

Contents lists available at ScienceDirect

Atmospheric Environment
journal homepage: www.elsevier.com/locate/atmosenv

Particulate and gaseous emissions from the combustion of different


biofuels in a pellet stove
E.D. Vicente a, M.A. Duarte a, L.A.C. Tarelho a, T.F. Nunes a, F. Amato b, X. Querol b,
C. Colombi c, V. Gianelle c, C.A. Alves a, *
a
b
c

Centre for Environmental and Marine Studies, Department of Environment and Planning, University of Aveiro, 3810-193 Aveiro, Portugal
Institute of Environmental Assessment and Water Research, IDAEA, Spanish Research Council (CSIC), C/Jordi Girona 18-26, 08034 Barcelona, Spain
Regional Centre for Air Quality Monitoring, Environmental Monitoring Sector ARPA Lombardia, 20129 Milan, Italy

h i g h l i g h t s
 The highest emission factors were obtained for agro-fuels.
 Organic carbon contributed no more than 30% of the PM10 mass.
 Mannosan and galactosan were not detected in almost all samples.
 Treated wood in pellets generated high contents of Pb, Zn and As in PM10.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 22 January 2015
Received in revised form
18 July 2015
Accepted 22 August 2015
Available online 28 August 2015

Seven fuels (four types of wood pellets and three agro-fuels) were tested in an automatic pellet stove
(9.5 kWth) in order to determine emission factors (EFs) of gaseous compounds, such as carbon monoxide
(CO), methane (CH4), formaldehyde (HCHO), and total organic carbon (TOC). Particulate matter (PM10)
EFs and the corresponding chemical compositions for each fuel were also obtained. Samples were
analysed for organic carbon (OC) and elemental carbon (EC), anhydrosugars and 57 chemical elements.
The fuel type clearly affected the gaseous and particulate emissions. The CO EFs ranged from 90.9 19.3
(pellets type IV) to 1480 125 mg MJ1 (olive pit). Wood pellets presented the lowest TOC emission
factor among all fuels. HCHO and CH4 EFs ranged from 1.01 0.11 to 36.9 6.3 mg MJ1 and from
0.23 0.03 to 28.7 5.7 mg MJ1, respectively. Olive pit was the fuel with highest emissions of these
volatile organic compounds. The PM10 EFs ranged from 26.6 3.14 to 169 23.6 mg MJ1. The lowest
PM10 emission factor was found for wood pellets type I (fuel with low ash content), whist the highest was
observed during the combustion of an agricultural fuel (olive pit). The OC content of PM10 ranged from
8 wt.% (pellets type III) to 29 wt.% (olive pit). Variable EC particle mass fractions, ranging from 3 wt.%
(olive pit) to 47 wt.% (shell of pine nuts), were also observed. The carbonaceous content of particulate
matter was lower than that reported previously during the combustion of several wood fuels in traditional woodstoves and replaces. Levoglucosan was the most abundant anhydrosugar, comprising 0.02
e3.03 wt.% of the particle mass. Mannosan and galactosan were not detected in almost all samples.
Elements represented 11e32 wt.% of the PM10 mass emitted, showing great variability depending on the
type of biofuel used.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Pellet stove
PM10
OC/EC
Anhydrosugars
Inorganic species

1. Introduction
Biomass combustion has been encouraged aiming at reducing
fossil fuel consumption. However, high emissions from incomplete

* Corresponding author.
E-mail address: celia.alves@ua.pt (C.A. Alves).
http://dx.doi.org/10.1016/j.atmosenv.2015.08.067
1352-2310/ 2015 Elsevier Ltd. All rights reserved.

fuel combustion in small-scale appliances like woodstoves and


replaces have been reported in several countries (Fine et al., 2004;
Gonalves et al., 2010; Schmidl et al., 2008). Other technologies are
available for domestic heating purposes with advantages from the
emission point of view. Small scale pellet heating systems are
installed in rising tendency. The wood pellet market has experienced a large growth in recent years as a result of the EU objective

16

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

to increase the share of renewable energy (Goh et al., 2012). The use
of pellets as fuel in small scale appliances for heating purposes has
been pointed out as suitable in order to reduce the emissions from
this sector (Pettersson et al., 2010). Between 2008 and 2010, the
production of wood pellets in EU increased by 20.5%, reaching 9.2
million tons in 2010, representing 61% of the global production. In
the same period, the EU wood pellet consumption increased by
43.5% to reach over 11.4 million tons in 2010, representing nearly
85% of the global wood pellet demand. In the segment of residential
heating, the main drivers for market expansion are often indirect
support measures for the installation of pellet stoves and boilers, as
well as the relative cost competitiveness of wood pellets compared
to traditional fuels, such as LPG heating oil and natural gas, especially in rural areas that are not yet supplied by gas grids (Cocchi
et al., 2011). For example, in Spain, the Ministry of Economy has
encourage the installation of boilers as part of the 2004e2012
Energy Plan, which aimed to promote the use of biomass, such as
pellets, olive pit and almond shell, as an energy source. In Portugal,
Spain and other southern European countries, the cork and olive oil
sectors generate large amounts of residues that can be used as fuel
for heating in small scale appliances (Garcia-Maraver et al., 2014).
In Portugal, the potential for producing pellets through the use of
agricultural residues is recognised. The energy potential derived
from almond residues, for example, was estimated to be
93 kton y1. Cereal straw and pruning residues are the agroresidues with more energy potential (Monteiro et al., 2012). Agricultural fuels present different physical and chemical characteristics compared with woody fuels (Verma et al., 2011; Obernberger
and Thek, 2004). For that reason the combustion of these fuels in
small scale appliances can be a challenge (Carvalho et al., 2008).
Pellet appliances are sensitive to variations of the ash-forming elements in the fuel due to slag formation on the burner, which is

problematic with respect to combustion efciency (Ohman


et al.,
2004), due to critical inorganic elements, such as alkaline metals,
that give rise to increased particulate emissions (Carvalho et al.,
2013). Ash accumulation and slag formation can even lead to shut
down of the appliance (Carvalho et al., 2008). The chemical
composition of the fuels has also inuence on the ash melting
behaviour, deposit formation and corrosion (Obernberger and
Thek, 2004).
Although good combustion conditions lead to lowest particulate
emissions, several studies have reported highest oxidative stress,
inammatory, cytotoxic and genotoxic activities and decreased
cellular metabolic activity from particles generated under efcient
combustion conditions rather than particles resulting from
smouldering combustion (Happo et al., 2013; Uski et al., 2014).
Besides the health effects, biomass combustion particles are efcient cloud condensation nuclei and can inuence the formation of
schl, 2005; Rose et al., 2010). In spite of the
precipitation (Po
increasing use of pellet stoves, their emissions are poorly typied.
The variability in properties of biomass is great and may signicantly inuence the efciency and environmental impacts associated with their use, constituting an issue of great research interest.
The aim of this research is to characterise the emissions
resulting from a Portuguese model of pellet stove with growing
market share in the residential sector, where different solid biofuels
(four types of pellets, olive pit, almond shell and shell of pine nuts)
have been burnt. The work comprises extensive quantitative and
qualitative data for gaseous compounds and particulate matter
(PM10). Particulate samples were analysed for organic (OC) and
elemental carbon (EC), anhydrosugars and 57 elements. A
comparative analysis of particulate phase emissions of parent
polycyclic aromatic hydrocarbons (PAHs) and their derivatives
(alkyl-PAHs, oxygenated-PAHs, azaarenes and nitrated-PAHs) from
the combustion of these biofuels in the pellet stove with those from

a manually red appliance can be found elsewhere (Vicente et al., in


press).
2. Experimental work
2.1. Combustion equipment, fuels and experimental procedure
The combustion experiments were carried out using a top-feed
pellet stove with a nominal output of 9.5 kW, manufactured in
Portugal by Solzaima, model Alpes (Fig. 1). The stove has an internal
pellet storage tank with 20 kg capacity and the fuel is supplied by
an auger screw to the top feed burner. The primary air is supplied
through holes in the bottom of the grate, and secondary air is feed
above the grate through three holes. Both primary and secondary
air are driven by an electric fan located downstream the combustion chamber. The primary air ow rate was monitored continuously during the combustion cycle using a mass ow meter. The
temperature was measured continuously using K-type thermocouples located at several points along the combustion and exhaust
system (combustion chamber, chimney and dilution tunnel). The
stove can be operated at ve levels of power output by automatically modifying the fuel feed rate and the exhaust gas fan speed. In
order to cover different behaviours by users, the emission factors
(EFs) of distinct gaseous components and particulate matter (PM10)
were determined for three levels of power output (lowest, medium
and highest). Since the differences in emissions between different
levels of power output for the same fuel were not statistically signicant, average EFs were obtained for each compound. The ignition of the fuel is made through an electrical resistance located on
the grate of the stove. The ignition phase was not included in the
results.
It should be pointed out that a short cleaning period of the grate
is programmed to occur. During the cleaning process, the fuel
supply decreases and the combustion air supply increases for a few
minutes in order to remove the bottom ashes from the grate; thus,
the lighter ash fractions are then carried out with ue gas.
Each of the combustion experiments was performed after a
preheating period of the pellet stove. When the level of power
output was changed, it took about 40 min to start the experiments
to ensure that the combustion process had already attained a new
steady operation condition.
The feeding rate for each fuel was evaluated by prior calibration
of the screw feeding system for the three levels of power output.
Also, in all experiments, pre-weighed fuel was poured into the fuel
hopper and after combustion the remaining fuel was re-weighed, in
order to verify the fuel consumption rate. The feed rate showed
great variations among fuels (Table 2). This can be related to the
differences in the physico-mechanical properties of the fuels, such
as the diameter and length, the bulk density, the ne content and
the mechanical durability (Carvalho et al., 2013; Verma et al., 2011).
To investigate the inuence of fuel quality on emissions, four
types of pellets were selected. These biofuels were selected based
on their commercial availability in order to represent the wide
range of pellets on the Portuguese market. Pellets type I were
commercial wood pellets made of golden wattle, cedar and pine.
Pellets type II were composed of 75% of lignocellulosic residues
and 25% of dust from the furniture manufacturing industry.
Pellets type III were composed of 65% of lignocellulosic residues
and 35% of dust from the furniture manufacturing industry.
Pellets type IV were made with a mixture of 50% of waste
woodchips (several woods from construction and demolition,
pine wood pallets, forest biomass, paper and paperboard) and
50% of dust from the furniture manufacturing industry. In addition, EFs for the combustion of olive pit, almond shell and shell of
pine nuts were obtained. This characterisation was carried out at

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

17

H
E

A
B
C

Fig. 1. Schematic representation of the experimental installation. A e Stove; B e Combustion chamber; C e Grate of the stove; D e Air ow meter; E e Exhaust duct (Chimney); F e
Gas sampling and analysis system; G e Water-cooled gas sampling probe; He Gas sampling pump; I e Gas condensation unit; J e Pitot tube; Ke Dilution tunnel; L,M e TECORA
PM10 sampling system; N e Fan.

the request of the Madrid City Council, because in the Autonomous Region these agro-fuels have shown a rapidly increasing
penetration in the residential heating sector due to the implementation of the national Energy Plan. Fuel properties, including
moisture, C, H, N, O, S and ash content, and also lower heating
value (LHV), are listed in Table 1. Moisture and ash content of
wood pellets for non-industrial use are two of the parameters

established by the European norm EN 14961-2. Pellets type III, IV,


olive pit and shell of pine nuts presented moisture content
higher than the limit of 10% established by the norm. The ash
content of pellets type II and III was higher than 3%, the
maximum allowed by the norm. The nitrogen content of pellets
type II, III and IV was also higher than the maximum value
allowed by the norm (1 wt. %). The content of major and trace

18

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

Table 1
Characteristics of the biofuels used in the experiments.

Proximate analysis (wt.%, as received)


Ultimate analysis (wt.%, dry basis)

Pellets type I

Pellets type II

Pellets type III

Pellets type IV

Olive pit

Shell of pine nuts

Almond shell

Moisture
Ash
C
H
N
S
O (by difference)

8.4
0.73
49.7
6.9
0.16
<0.01
42.5
18.3

8.8
3.2
47.4
6.58
2.31
<0.01
40.5
17.5

10.9
3.8
48.3
6.53
2.06
<0.01
39.3
17.4

10.7
2.0
47.4
6.79
2.11
<0.01
41.7
17.8

12.9
0.66
50.9
6.59
0.21
<0.01
41.6
18.5

12.9
1.3
49.8
6.59
0.30
<0.01
42.0
18.7

9.5
1.4
49.3
6.76
0.34
<0.01
42.
18.4

Al2O3
CaO
Fe2O3
K2O
MgO
Na2O
P2O5
SO3
Zn
As
Pb
Cr
Cu

0.02
0.2
0.02
0.1
0.03
0.01
0.01
0.03
92
1.8
2.3
4.2
1.5

0.1
0.7
0.1
0.1
0.1
0.1
0.04
0.2
181
3.8
119
4.9
10

0.1
0.7
0.2
0.1
0.1
0.1
0.04
0.2
304
7.3
207
7
17

0.1
0.3
0.1
0.1
0.1
0.1
0.1
0.2
76
1.9
19
4.8
7.1

<0.01
0.1
0.01
0.3
0.02
<0.01
0.03
0.03
3.90
16
4.8
<0.5
5.0

0.04
0.2
0.1
0.2
0.1
0.02
0.03
0.1
31
<0.5
8.6
2.6
8.7

0.03
0.3
0.1
0.3
0.1
0.1
0.1
0.05
57
1
4.4
3.7
8.6

LHV (MJ kg1)


wt.% oxide

mg kg1 of dry fuel

Table 2
Operating conditions, PM10 emission factors and carbonaceous content of particulate matter.
Fuel

Number of
experiments

Fuel feed
rate (kg h1)

Pellets type I
Pellets type II
Pellets type III
Pellets type IV
Olive pit
Shell of pine nuts
Almond shell

4
9
7
6
6
8
6

1.18
1.40
1.12
1.44
0.88
0.96
1.36

0.17
0.23
0.17
0.10
0.07
0.18
0.29

Dilution
ratio
35.2
29.9
26.0
29.7
31.0
30.4
25.4

9.2
3.4
3.4
6.6
4.9
12.1
3.1

O2 concentration (% v, dry
gases) in the exit ue gas
16.9
16.6
18.2
15.7
17.0
18.4
17.2

1.3
0.39
0.26
0.53
0.33
0.17
0.42

elements in the fuels were determined by Inductively-Coupled


Plasma Atomic-Emission Spectrometry (ICP-AES) and by
Inductively-Coupled Plasma Mass Spectrometry (ICP-MS),
respectively, following the same methodology described in Section 2.3 for the particulate matter lters. Pellets (made up of
wood and wood wastes) are characterised by high Ca and K, due
to the occurrence of wewellite, K salts and organic-bearing K. The
use of wood preservatives, such as chromated copper arsenate
(CCA), give rise to relatively high and soluble concentrations of
Cr, Cu and As. Olive pit, almond shell and shell of pine nuts show
higher K than Ca concentrations and relatively low content of
most trace elements with the exception of As in olive pit
(Table 1).
2.2. Flue gas measurements
The combustion ue gas was sampled at the exit of the chimney
by means of a heated (at 180  C) sampling line and conducted to an
online Fourier Transform Infrared Gas analyser (FTIR Gasmet,
CX4000). This equipment has a multicomponent measurement
capability, which enables, among others, the real-time monitoring
of CO, TOC, CH4 and HCHO. TOC includes CH4, ethane (C2H6), ethane
(C2H4), propane (C3H8), hexane (C6H14) and HCHO. The sample cell
heating ensures that high water vapour concentrations or corrosive
gases will not pose a problem. Since the water content of the
sample gas is continuously measured with the Gasmet, the results can be reported on either a wet or dry basis. The concentration of several gaseous compounds was recorded as wet gas,
and then corrected to dry basis based on the H2O concentration
measured.

Flue gas temperature ( C) in


the combustion chamber
683
681
743
800
828
623
876

48.9
42.9
10.1
3.10
57.2
48.8
34.4

PM10
(mg MJ1)
26.6
86.4
102
75.6
169
117
112

3.14
13.6
8.63
9.39
23.6
33.9
4.05

OC
(mg MJ1)
7.40
14.3
8.38
15.2
48.7
18.7
12.1

1.86
8.46
1.21
6.28
30.9
3.12
3.00

EC
(mg MJ1)

OC/
EC

4.2
1.0
0.9
1.7
9.4
0.3
1.7

1.77
13.9
9.67
8.89
5.18
54.5
7.00

0.44
8.32
2.81
2.07
3.51
23.5
0.94

2.3. Particulate matter sampling and analysis


Particulate matter (PM10) was collected after the dilution of the
exhaust with atmospheric air in a tunnel under isokinetic conditions. The sampling point was located 10 m downstream the dilution tunnel entrance. The dilution system was described in previous
studies (Alves et al., 2011; Gonalves et al., 2010; Calvo et al., 2014).
The dilution ratio (DR) was dened as the ratio between the total
diluted ow (in the dilution tunnel) and the undiluted ue gas ow
(in the chimney) and ranged from 25.4 to 35.2 (Table 2). The
sampling train included a PM10 inlet head, a pump, and a control
and data acquisition system, all part of a TCR TECORA (model
2.004.01) instrument. The equipment has been operated at a ow
of 2.3 m3 h1. The particulate matter sampling was made under
steady state operating conditions, which have been evaluated by
continuous monitoring of ue gas composition. A cascade impactor
TCR Tecora (model MSSI) was used to obtain size segregated particulate matter samples. For each fuel, sampling with cascade
impactor was carried out at each power of level output. These
samples were collected in parallel with the PM10 lters that have
been used to perform the chemical characterisation reported in the
present study. The multistage cascade impactor, designed according to ISO23210-2009 norm for the nozzle dimensioning, allows
separation into 3 particulate size fractions: >10 mm, between 2.5
and 10 mm and below 2.5 mm. The sampling ow rate was 3 m3 h1.
The particulate matter (PM10) samples for gravimetric and chemical
analyses were collected on 47 mm diameter quartz bre lters prebaked at 500  C for 6 h, to remove organic contaminants. The lters
were kept in a desiccator to stabilise without hydration or
contamination. The gravimetric quantication was performed with

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

a microbalance (RADWAG 5/2Y/F with an accuracy of 1 mg). Filter


weight was obtained from the average of six measurements, when
the variations were less than 0.02%.
The organic (OC) and elemental carbon (EC) content of PM10 was
analysed by a thermal optical transmission technique. This method
allows the differentiation of various particulate carbon fractions
through volatilisation and oxidation to CO2 under controlled
heating. The CO2 produced is then quantied in a non-dispersive
infrared (NDIR) analyser. The blackening of the lter is monitored
using a laser beam and a photodetector, which enables separating
the EC formed by pyrolysis. A more detailed description of the
method can be found elsewhere (Gonalves et al., 2014).
The analysis of major and trace elements was performed by ICPAES and ICP-MS. For each biofuel, punches from the various replicate lter samples obtained for the three levels of power output
were combined and analysed together in order to obtain mean
values for each combustion condition. As proposed by Querol et al.
(2001), each set of lter punches was subjected to acid digestion
(1.25 mL HNO3: 2.5 mL HF: 1.25 mL HClO4). Three multi-elemental
solutions Spec 1 (rare earth elements, REE), Spec 2 (alkalis, earth
alkalis, and metals) and Spec 4 (Nb) were used to construct
external calibration curves. The mean precision and accuracy were
controlled by repeated analysis of 0.025 mg of NBS-1633b (y ash)
reference material (NIST, Gaithersburg, MD, USA). They fall below
typical analytical errors and are in the ranges of 3e5% and <10%, for
ICP-AES and ICP-MS, respectively. The detection limits were 0.01 ng
m3 for most of the trace elements analysed.
In order to quantify levoglucosan and its isomers, punches of
various replicate lter samples obtained for each level of power
output were extracted with milliQ water during 1 h in an ultrasonic
bath. Determinations of levoglucosan were performed using a
Metrohm 881.0020 Compact IC Pro ion cromatographic system
equipped with the amperometric 896.0030 Professional Detector
Metrohm (Metrohm IC Application Work, 2008, 2009). The Metrosep Carb 1 e 150/4.0 column was used, with another column
ASupp15 above, allowing a better separation of mannosan and
galactosan. The columns are thermostated at 40  C. As eluent it was
used, NaOH 70 mM, prior outgassing before the application in order
to remove CO2. The elution ow was 0.6 mL min1. The system was
calibrated using a solid standard of levoglucosan (ACROS ORGANICS, >99%) diluted in milliQ water. Different concentrations of
standard, from 0.01 to 10 ppm, were prepared.
3. Results and discussion
3.1. Gaseous emissions
Gaseous emissions are greatly affected by the biofuel type
(Fig. 2). Excepting pellets type I, higher gaseous emissions for
agricultural fuels than for wood pellets were observed. Ashes of
pellets type I presented a much higher C content (27%) compared
with that of pellets type II, III, and IV (1.5e3.7%). The t-test
(a 0.05) revealed that the difference in the CO EFs from pellets
type I and the other three types of wood pellets is statistically
signicant (p < 0.0001). On the other hand, the differences between
emissions from pellets type I and two of the three agricultural fuels
were found to be statistically insignicant (p 0.9631 and
p 0.1883 for shell of pine nuts and almond shell, respectively).
The CO EFs ranged from 90.9 19.3 (pellets type IV) to
1480 125 mg MJ1 (olive pit). The high CO EFs observed for
agricultural fuels indicate poor combustion conditions. The CO EFs
obtained in the present study for agricultural fuels are in the range
of values obtained for conventional batch combustion
(830e2700 mg MJ1) (Lamberg et al., 2011a). Wood pellets present
lower bulk densities than agricultural fuels allowing favourable air

19

excess in the combustion chamber, since lower density fuel delivers


less energy into the combustion chamber (Verma et al., 2012). The
CO EF was about 16 higher for olive pit than for pellets type IV.
Carvalho et al. (2013) reported eleven to fourteen times higher CO
emissions from several agricultural fuels than from wood pellets.
Although the difference between agricultural and wood fuels is
close to the difference documented by Carvalho et al. (2013), the CO
EFs are much higher in the present study (Table 3). The differences
between the results of the present study and literature data can be
associated with the pelletisation of the agricultural fuels, which can
contribute to improved feeding into the stove and, consequently, to
more favourable operating conditions. Fuel blending can potentially contribute to reduce emissions from the combustion of agricultural fuels (Miranda et al., 2012). In the present study, the
chemical composition of the fuel was found to have inuence on
the CO EFs. The CO EFs were negatively correlated with the AlO3
(R2 0.88) and the SO3 (R2 0.79) and positively correlated with
the K2O (R2 0.71) contents in the fuels.
The difference in the gaseous emissions between agricultural
and wood pellets can be related to the formation of ash agglomerates in the grate of the stove, creating an irregular air combustion
ow rate through the fuel bed, which has a negative effect in the
oxidation of the gaseous compounds. The difference between olive
pit CO EF and wood pellets (type I, II, III and IV) is statistically signicant (p < 0.0001). The difference in the CO EFs from almond
shell and shell of pine nuts is not statistically signicant
(p 0.2302), in contrast to the statistically signicant difference in
the EFs from olive pit and the other two agricultural fuels
(p 0.0001 and p 0.0003 for shell of pine nuts and almond shell,
respectively). The higher moisture content of the agricultural fuels
may also have led to poor combustion conditions and consequently
to higher CO EFs. Verma et al. (2012) used a multi-fuel boiler
(40 kW) to test peat pellets, apple pellets, wheat straw pellets with
hydrated lime as additive and sun ower husk pellets. The authors
observed that the CO emissions for the combustion of peat pellets
were 50.0, 6.1 and 2.5 times higher than those of apple, straw and
sun ower husk, respectively. The results obtained in the present
and previous studies highlight the importance of the fuel in
continuous combustion appliances. These devices are designed for
wood pellets with a specic size and shape and for that reason the
feeding system and combustion chamber may not be prepared for
alternative fuels. Such fuels are chemically different than wood
pellets presenting higher concentrations of ash forming elements
and therefore causing operating problem that maximise CO emissions concomitantly with slagging on the grate of the stove.
Carvalho et al. (2013) investigated the performance of a pellet
boiler red with agricultural fuels and found that the combustion
chamber was not optimal for these fuels. Furthermore, the authors
also reported that the boiler was incapable to adjust an optimal
excess air ratio and the efciency was gradually reduced due to ash
accumulation on the heat exchangers in consecutive combustion
experiments.
With respect to TOC (Fig. 2), wood pellets generated lower EFs
than other fuels. The emission values ranged from 2.2 0.2 (pellets
type I) to 68.7 11.4 (olive pit) mg MJ1. Olive pit was the fuel with
highest EF, which is possibly related to the fact that olive fruit has a
stony pit containing 30% lipids and 20% carbohydrates. Johansson
et al. (2004) compared the emissions from different types of
combustion appliances. The authors observed that the operation
and combustion appliance exert a great inuence on TOC emissions. These variations are reected in the discrepancies between
results. In our study, it was observed that the fuel characteristics
also play a major role in TOC emissions. As observed by Johansson
et al. (2004), the TOC emissions correlated with the CO emissions
(R2 0.86). Similar to CO EFs, the difference in the TOC EFs from

20

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

Fig. 2. CO, TOC, HCHO and CH4 emission factors for the combustion of the distinct biofuels.

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

Fuel

CO EF (mg MJ1)

Reference

Pellet stove (8 kW)


Pellet boiler (25 kW)
Pellet stove (8 kW)

Low-quality cheap pellets and high quality pellets with DIN-PLUS certication

88
350
101e201
207e355
3.55e335a
280
224
223
90.9e740
732e1480

Ozgen et al. (2014)

Pellet boiler (25 kW)


Pellet boiler (15 kW)

Pellet stove (9.5 kW)


a

Commercial pine stem and spruce stem pellets


Alder, birch, and willow stem pellets
Pine steam wood pellets
Hay pellets
Wheat bran pellets
Straw pellets
Wood pellets
Agro fuels

Sippula et al. (2007)


Lamberg et al. (2011b)
Carvalho et al. (2013)

This study

Different operating conditions (boiler load, primary and secondary air supplies).

pellets type I and the other three types of wood pellets is statistically signicant (p < 0.0001), as well as the difference between the
TOC EFs for olive pit and wood pellets (p < 0.0001). The higher
emissions of TOC from agricultural fuels are, at least in part, due to
fuel characteristics, which impair the combustion efciency, as
discussed above for CO. Differences statistically insignicant were
found when comparing the TOC EFs from pellets type I and almond
shell (p 0.1981), pellets type IV and pellets type II (p 0.1190) and
almond shell and shell of pine nuts (p 0.2664).

Formaldehyde (HCHO) is known to induce severe poisoning,


irritation and other immune toxic effects (Tang et al., 2009). The
meta-analysis by Zhang et al. (2009) support an association between formaldehyde and leukemia. Cerqueira et al. (2013) conducted experiments using ve different types of common wood
species growing in Portugal in order to characterise formaldehyde
and acetaldehyde emissions from residential combustion in a
woodstove. Assuming a heating value of 18 MJ kg1 of dry biofuel
(Telmo et al., 2010), EFs for formaldehyde were estimated to be

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

between 36.3 and 98.4 mg MJ1. In the present study, the formaldehyde EFs ranged from 1.01 0.1 (pellets type II) to 36.9 6.3
(olive pit) mg MJ1 (Fig. 2). Formaldehyde emissions have shown to
depend on the combustion appliance (Tissari et al., 2007). In the
present study, it was observed that the type of fuel also plays a
major role on the emissions of this compound. Formaldehyde is an
important product in biomass pyrolysis at temperatures higher
than around 200  C (Zhang et al., 2011). The differences between
formaldehyde EFs may be due to dissimilar cellulose and lignin
lez et al., 2009). Furthermore, variable
contents in the fuels (Gonza
combustion efciencies also inuence these emissions (Yokelson
et al., 1997). The difference in the HCHO EFs from pellets type I
and the other three types of wood pellets is statistically signicant
(p < 0.0001), as well as the difference between the HCHO EF for
olive pit and the other fuels. No signicant differences were found
between the EFs from almond shell and shell of pine nuts
(p 0.2067) or between pellets type I and shell of pine nuts
(p 0.6324). The HCHO EFs correlate linearly with the TOC EFs
(R2 0.97) and with CO EFs (R2 0.84).
One of the most important volatile organic compound (VOC)
from biomass combustion is methane (CH4), which is an effective
greenhouse gas (Obaidullah et al., 2012). Johansson et al. (2004)
compared the emissions from commercial residential boilers red
with wood logs and wood pellets. In their study, it was observed
that, in all cases, methane was the VOC with highest concentration.
Pettersson et al. (2011) characterised the emissions from a woodstove with variations in fuel, appliance and operating conditions.
The authors also found that methane was the predominant VOC
with EFs ranging from 9 to 1600 mg MJ1. Tissari et al. (2008a)
characterised the effect of two different combustion conditions
on particulate and gaseous emissions from a conventional masonry
heater, namely normal combustion and smouldering combustion.
The authors found that the emission of methane from smouldering
combustion was 11-fold those from normal combustion. Thus, CH4
emissions strongly depend on the combustion conditions. Olsson
llstrand (2006) studied the emissions of organic comand Kja
pounds from wood burning in a modern eco labelled residential
boiler (30 kW). The authors veried that CH4 emissions decrease
with increasing combustion efciency. Leskinen et al. (2014) also
observed differences among the CH4 EFs from distinct combustion
conditions, reporting average values of 0.3, 4.3 and 115 mg MJ1 for
efcient, intermediate and smouldering combustion, respectively.
In the present work, the CH4 EFs ranged from 0.23 0.03 (pellets
type V) to 28.7 5.7 (olive pit) mg MJ1 (Fig. 2). The statistical
analysis revealed that the difference between the olive pit CH4 EF
and all the other fuels is statistically signicant (p < 0.0001), as well
as the difference between the pellets type I EF and the other wood
pellets (p < 0.0001). A clear linear correlation was found between
the CH4 and TOC EFs (R2 0.99) and CO EFs (R2 0.87).
3.2. PM10 emission factors
Particulate matter emissions from residential wood combustion
for heating purposes is one topic of great concern. With the
exception of pellets type II and shell of pine nuts, the sizesegregated samples indicated that a mass fraction higher than
80% of the total particulate matter emitted is concentrated in the
nest range. The very ne particles in smoke can go deep into the
lungs. The technical characteristics of residential combustion appliances play a major role on emissions. Appliances may be divided
into two groups: manually fed batch-combustion systems and
automatically fed systems. The fuel properties may also play a
signicant inuence on the amount and composition of particle
emissions (Fernandes et al., 2011; Schmidl et al., 2008; Lamberg
et al., 2011a,b; Sippula et al., 2007). The transient combustion

21

phases were not included in the results of the present study.


However, it is important to take into account that the unsteady
phases can signicantly affect the emissions. Toscano et al. (2014)
reported that unsteady state phases, such as the startup phase,
can signicantly affect the emission factors. In their study, the PM
concentration was up to 72% more than those measured in steady
state condition.
PM10 EFs from the combustion experiments are listed in Table 2.
The EFs ranged from 26.6 3.14 (pellets type I) to
169 23.6 mg MJ1 (olive pit). The difference in the PM10 EFs from
pellets type I and the other three types of wood pellets is statistically signicant (p < 0.0001), as well as the difference between
PM10 EFs for olive pit and wood pellets (p < 0.0001). In general,
higher EFs were obtained for biomass fuels other than pellets,
which may be related to the fuel physical characteristics, such as
dimensions (length and diameter), ne content and particle density. These properties are important as regards the fuel supply to
the burner and affect the combustion behaviour (Carvalho et al.,
2013; Obernberger and Thek, 2004; Verma et al., 2011), which, in
turn, inuence the equipment performance and its emissions
(Table 2). The pelletisation of agricultural fuels and the use of
blended pellet fuels can contribute to optimise the combustion
process and subsequently reduce the emissions (Mediavilla et al.,
2009; Heschel et al., 1999). Insignicant differences were only
found between the PM10 EF from pellets type II and IV (p 0.1166)
and from almond shell and shell of pine nuts (p 0.7282). The
chemical composition of the fuel can signicantly inuence the
particulate emissions (e.g. Obernberger and Thek, 2004). Inorganic
aerosols formation occurs through nucleation and condensation of
the ash forming vapours like alkali metals and easily volatile heavy
metals. These compounds are released during combustion from the
fuel to the gas phase. They have a high vapour pressure and, as soon
as the vapour pressure exceeds the saturation pressure, particles
are formed by nucleation or they condense on surfaces of existing
particles (Skotland, 2009; Ingwald and Gerold, 2006). The high EFs
observed for the agricultural biomass fuels may be related to the
fact that these fuels present higher K content than woody pellets
(Table 1). As observed in previous studies (e.g. Sippula et al., 2007),
in the present study the PM10 EFs seems to correlate with the fuel
K2O content (R2 0.57). Excluding pellets type I, a clear linear
correlation was found between the PM10 EFs from the other fuels
and the CO EF (R2 0.91), the CH4 EFs (R2 0.95), the HCHO EFs
(R2 0.99) and the TOC EFs (R2 0.94), indicating that high particulate emissions are associated with high emissions of products
from incomplete combustion. Although Southern European countries have not yet imposed standards, stringent particulate matter
emission limits for solid fuel heating appliances are currently being
discussed in other countries. For example, in Germany, from
January 2015 onwards, stoves have to comply with more stringent
emission limit values, ranging between 13 and 27 mg MJ1,
depending on the type of fuel. Except for pellets type I, the emissions from all other biofuels far exceed the limits set by the Germans. However, the PM10 EFs observed in the present study are
8e10 times lower than those obtained in a traditional replace
(Alves et al., 2011; Fernandes et al., 2011). Fig. 4 presents a comparison between the results obtained in previous studies using
distinct appliances and the results recorded in the present work.
Open replaces present the highest particulate EFs. The low temperatures in these appliances contribute to very inefcient combustions. Smouldering is the prevalent combustion phase. The
continuous improvement of combustion technologies has
contributed to dramatic reductions in emissions.
Fernandes et al. (2011) found PM2.5 EFs of 699 327, 447 169
and 103 50.6 mg MJ1 for the combustion in a replace, woodstove and eco-labelled woodstove, respectively (Fig. 3). Although

22

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

the emissions resulting from the combustion in automatically red


appliances are signicantly lower than those from log woodstoves
and replaces, variations can be observed. These variations may be
associated either with the fuel properties or with the burning mode
(full or part load operation) (Schmidl et al., 2008; Garcia-Maraver
et al., 2014; Riva et al., 2011; Lamberg et al., 2011b). GarciaMaraver et al. (2014) found that gaseous and PM emissions are
signicantly affected by the fuel type rather than the boiler thermal
load. Ozgen et al. (2014) tested two types of pellets (low quality
cheap pellets and high quality pellets with DIN-PLUS certication)
in a pellet stove (8 kW) and boiler (25 kW). The authors reported an
average emission factor of 109 mg MJ1 for the pellet stove and of
61 mg MJ1 for the pellet boiler.
Automatically red appliances allow operating under more
stable and efcient conditions. Particulate matter resulting from
the combustion in these appliances is mainly composed of alkali
salts. The fuel ash content and composition have a great inuence
on emissions. Thus, the variability in the PM10 EFs observed in the
present study for the different fuels may be explained by the
different ash contents. Pellets type I showed the lowest emission
factor and the lowest ash content (0.73 wt.%), whilst pellets type III
exhibited the highest emission factor and ash content (3.8 wt.%).
Sippula et al. (2007) found that the fuel ash content correlated
linearly with the PM1 emission.
3.3. PM10 chemical composition
Particles resulting from residential wood combustion are
composed of soot, organic matter and ne y ash (Obaidullah et al.,
2012; Orasche et al., 2012; Blling et al., 2009; Tissari, 2008). Particles from manually operated small appliances are, in general,
dominated by organic matter since the combustion conditions are
more incomplete. Table 2 displays the EFs of carbonaceous constituents for the distinct combustion experiments. The OC EFs
ranged from 7.40 1.86 (pellets type I) to 48.7 30.9 (olive pit)
mg MJ1. The differences between the OC EFs from pellets type I
and pellets type II and III are not statistically signicant (p 0.1430
and p 0.3118, respectively), and neither are the differences

Fig. 3. Comparison between particulate emission factors (mg MJ1) from different
types of small-scale combustion appliances and the present study.

P
Fig. 4. Chemical composition of PM10 ( chemical constituents OC  1.8 EC
Oxides), for the combustion of the distinct biofuels.

between pellets type II, III and IV. Statistically signicant differences were observed between the EF from pellets type I and IV
(p 0.0451). The difference between the olive pit OC EF and all the
other fuels was found to be statistically signicant. The EC EFs
ranged from 1.77 0.44 (pellets type I) to 54.5 23.5 (shell of pine
nuts). The shell of pine nuts EC EF is signicantly higher than the EC
EF from the other agricultural fuels (p 0.0003 and p 0.0004, for
olive pit and almond shell, respectively). The EC EF from pellets
type I was statistically different from almost all the other fuels. No
statistically signicant difference was found between the pellets
type I and olive pit EFs (p 0.0947).
Total carbon represented from 17.0 (almond shell) to 62.6 (shell
of pine nuts) wt.% of the PM10 mass. The OC EFs were found to
correlate with the TOC EFs (R2 0.84), the HCHO EFs (R2 0.78),
and the CH4 EFs (R2 0.82). Although the carbonaceous content of
particulate matter in the present study was lower than that obtained during the combustion in traditional woodstoves and replaces (Alves et al., 2011), high OC and EC emissions were observed
during the combustion of several fuels. The EC content in samples
from the combustion of shell of pine nuts was much higher than
that in PM10 from other fuels. Primary soot particles are formed
mainly in the ame from hydrocarbons. Depending on the temperature and oxidative conditions, primary particles may be burned
or remain in the particulate phase. As a consequence of the insufcient mixing of combustion gases and air, the ame zone always
contains fuel-rich areas providing the conditions for the formation
of soot particles (Torvela et al., 2014; Obaidullah et al., 2012; Tissari,
2008). This mechanism may be the explanation for the high EC
content in the particulate matter from the combustion of shell of
pine nuts. The difference in the EC content between emissions from
shell of pine nuts and the other two agricultural fuels may also be
related to the alkali salts content of the fuels. Almond shell and
olive pit contain much more Na and K. Since alkali metals can have
a catalytic effect in the soot oxidation process and thereby increase
the soot burnout rate (Wiinikka, 2015), this may be the reason for
the lower EC EFs compared with shell of pine nuts. It has been
described that the composition of the fuel burned plays an
important role on soot formation (Baeza-Romero et al., 2010). From
modelling results, Fitzpatrick et al. (2008) reported that the amount
of soot is dependent on the cellulose/lignin ratio. One of the
decomposition products of lignin is eugenol, which has been found
to greatly contribute to soot formation during wood combustion
(Fitzpatrick et al., 2008). The combustion of shell of pine nuts

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

generated higher amounts of EC in the particulate matter than OC,


whereas the opposite was observed for olive pit. It should be
recalled that this latter fuel was also the one with highest CO and
TOC EFs.
Improved combustion efciency with higher combustion temperatures and aming combustion contribute to higher EC emissions than those resulting from the combustion in appliances with
lower efciency like replaces and traditional woodstoves
(Fernandes et al., 2011). Gonalves et al. (2010) tested several fuels
in an eco-labelled log woodstove (6 kW) and found that total carbon represented 43.9e63.2 wt.% of the PM2.5 mass. Schmidl et al.
(2011) observed higher total carbon content in the particulate
mass emitted from traditional manually red appliances with EC
contributing to about 30% and OC to 30e40% of the PM mass.
Leskinen et al. (2014) studied the physico-chemical properties of
particles emitted under three different combustion conditions
(efcient, intermediate and smouldering combustion) and reported
that, during efcient combustion, particles were composed of ashrelated material, while during smouldering conditions, they contained more EC than OC. The PM1 from intermediate conditions
consisted mainly of OC and EC in almost equal fractions.
In the present study, the OC/EC ratio ranged from 0.9 to 4.2 and
from 0.3 to 9.4 for pellets and agro fuels combustion, respectively
(Table 2). These values are within the ranges reported in the literature (Table 4), since combustion parameters such as fuel type,
temperatures in the rebox, fuel feed rate, etc., are believed to
signicantly inuence the emission composition. The burning
conditions have great inuence on the OC to EC ratios, as can be
seen in Table 4. The types of fuel burned, as well as the kind of
appliance, are critical factors affecting the carbonaceous emissions.
Zhang et al. (2013) have evaluated prescribed burning versus
controlled burning in a stove. The authors reported lower OC/EC
ratios when burning wooden logs compared to combustion of
greener fuel types and observed lower OC/EC ratios in woodstove
burns compared to prescribed burns (Table 3). Gonalves et al.
(2010) reported OC to EC ratios in the range of 0.9e4.4 in particles emitted from an eco-labelled woodstove and different types of
Portuguese woods (Table 3).
In order to obtain a more precise mass balance, the OC mass was
converted to total mass of organic matter (OM) using a factor that
accounts for oxygen, hydrogen, and some other atoms present in
the organic material. The adopted OM/OC ratio was 1.8, which is
typical for wood smoke aerosols (Turpin and Lim, 2001). This
conversion allows to reconstruct the PM10 mass taking into account
the gravimetric data and the mass of all quantied species. The OM
fraction varied from 14.8 (pellets type III) to 51.9 (olive pit) wt.% of
the PM10 mass (Fig. 4).
As displayed in Table 5, levoglucosan, mannosan, and galactosan contents in particulate matter samples showed variations,

23

probably due to differences in the relative content of cellulose and


hemicellulose in diverse biofuels tested in the present study
(Engling et al., 2006). Levoglucosan was the most abundant
anhydrosugar, encompassing 0.02e3.03 wt.% of the particle mass.
This sugar was detected in almost all samples. For pellets type II
and III, levoglucosan was only detected during the operation at the
lower level of power output. Mannosan and galactosan were absent from almost all samples. The yields of mannosan and galactosan decreased more rapidly than those of levoglucosan under
increasing combustion severity (temperature and duration) (Kuo
et al., 2011).
Levoglucosan, mannosan and galactosan have been used as
tracers for residential wood combustion and wildres (Fine et al.,
2004; Engling et al., 2006; Vicente et al., 2012). The levoglucosan
to mannosan ratio has been described to be wood type specic,
with low ratios for softwoods and higher values for hardwoods
(Engling et al., 2006; Schmidl et al., 2008). However, the use of
levoglucosan as a quantitative tracer may be associated
with large uncertainty since laboratory measurements of levoglucosan emissions have shown large variations depending on
the type of stove, biofuel quality, and operator's behaviour
(Hedberg et al., 2006). Schmidl et al. (2011) studied the inuence
of system (manually and automatically red appliances) and
wood type on anhydrosugars emissions. The authors observed
that automatically red appliances did not emit detectable
amounts of anhydrosugars in full- and part-load operation. Levoglucosan and mannosan were only detected during the start-up
phase.
Recent studies of levoglucosan suggest that the transglycosylation process (cellulose degradation pathway) occurs at
lower temperatures than previously assumed, between 150 and
350  C (Kuo et al., 2008), with maximum yields at 250  C, regardless
of plant species. Therefore, levoglucosan is not a suitable tracer for
sophisticated appliances with automatically red wood combustion in which high temperatures are reached.
Inorganic elements accounted for about 10.6e31.7 wt.% of the
PM10 mass (Table 6). The dominant inorganic species in the PM10
samples from pellets type I and II, olive pit, and almond shell were
potassium and sulphate. The combustion of pellets type III and IV
and shell of pine nuts generated higher amounts of phosphorous. A
high phosphorous content in particulate matter is usually typical
for cereal fuel combustion (Tissari et al., 2008b). The proportion of
potassium in PM10 inorganic fraction was 66.0% for pellets type I,
40.6% for pellets type II, 86.8 for olive pit and 88.4% for almond
shell. It was not detected in samples from pellets type II and III. In
addition to these compounds, almost all samples contained zinc,
lead, magnesium and sodium and, to a lesser extent, other elements. For pellets type II, III and IV, high contents of zinc
(14.6e19.9 wt.% of total inorganic mass), lead (7.6e17.6 wt.% of total

Table 4
Comparison between OC/EC ratios from this study and literature values.
Appliance

Fuel

Fraction sampled

OC/EC

Reference

Fireplace
Woodstove

Portuguese woods and briquettes


Wooden logs (mainly pine)
Leaves/duff (Green foliage and branches from two
dominant shrubs: manzanita and bitterbrush)
Spruce wood logs
Portuguese woods
Spruce pellets
Spruce pellets
Wood pellets
Agro fuels

PM2.5
PM2.5

2.2e35.5
2.8 1.3
12.1 3.3

Gonalves et al. (2011)


Zhang et al. (2013)

PMtot
PM2.5
PMtot

11.5ae1.2b
0.9e4.4
4.4ae5.9b
8.5ae71.0b
0.9e4.2
0.3e9.4

Orasche et al. (2012)


Gonalves et al. (2010)
Orasche et al. (2012)

Log woodstove (8 kW)


Log woodstove (6 kW)
Pellet stove (13 kW)
Pellet boiler (25 kW)
Pellet stove (9.5 kW)
a
b

Cold start.
Nominal load.

PM10

This study

24

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

Table 5
Weight percentage (wt.%) of anhydrosugars in PM10.
Fuel

Level of power output

Levoglucosan

Mannosan

Galactosan

Pellets type I

Lowest
Medium
Highest
Lowest
Medium
Highest
Lowest
Medium
Highest
Lowest
Medium
Highest
Lowest
Medium
Highest
Lowest
Medium
Highest
Lowest
Medium
Highest

3.03
1.39
2.88
0.14
<ld
<ld
0.02
<ld
<ld
0.25
0.23
0.16
0.41
0.27
0.05
0.27
0.14
0.43
0.37
<ld
0.11

<ld
0.44
1.55
0.07
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
0.31
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld

<ld
0.37
0.73
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld

Pellets type II

Pellets type III

Pellets type IV

Olive pit

Shell of pine nuts

Almond shell

ld e limit of detection (0.075 mg cm2).

inorganic mass) and arsenic (0.079e0.030 wt.% of total inorganic


mass) were recorded. The use of CCA-treated wood in pellets has
been already reported to be problematic from the point of view of
heavy metals emissions (Wasson et al., 2005). The high concentrations of Pb in samples could be due to Pb-based paint on old
wood or uptake of Pb from soil contaminated by lead arsenate
pesticide (Sander, 1997).
Leskinen et al. (2014) reported particulate mass contents of Zn
and K of 5.5, 1.0 and 0.4 and 28.0, 6.7 and 1.5% for efcient, intermediate and smouldering combustion, respectively. The authors
found that the ash emissions were mainly composed of potassium,
sulphate and chloride at almost equal concentrations. However, the
fraction of sulphate was higher in samples from efcient combustion than in the other two conditions.
Taking into account that elements exist at the highest oxidation
states, it is recommended to convert the measured element concentrations into the respective mass concentrations of the most
common oxides for PM10 mass balances. When accounting for the
unmeasured oxygen, elements in their oxide form (e.g. Al2O3, CaO,
Fe2O3, K2O, MgO, Na2O, P2O5, etc.) represented 21.6 (shell of pine
nuts) to 42.2 (almond shell) wt.% (Fig. 4) of PM10.
4. Conclusions
This paper constitutes an attempt to quantify particle and
gaseous emissions from the combustion in a pellet stove of
different biofuels. It was found that the fuel type and its properties
can signicantly inuence the gaseous and particulate EFs from
residential combustion appliances. In addition, through the
organic and elemental carbon mass fractions, it was possible to
verify that the chemical composition of particles can also vary
noticeably. Combustion in modern appliances can signicantly
reduce PM emissions when compared to old type wood combustion systems.
The highest EFs were obtained for biomass fuels other than
pellets. The ash accumulation and slag formation on the grate of
the stove, as well as the physical characteristics of the agricultural fuels that difcult the feeding to the burner, impair the
combustion conditions and generate high gaseous and particulate emissions. The pelletisation of agricultural fuels and the use
of blended pellet fuels can contribute to optimise the combustion

process and subsequently reduce the emissions. It was possible


to verify that the enhanced emissions of CO were accompanied
by emissions of other unoxidised components like TOC, HCHO
and CH4 during agro-fuels combustion. Olive pit presented the
highest EFs for these compounds. CH4, which have much higher
global warming potential than CO2, accounted for a signicant
fraction of TOC in all experiments. The lowest PM10 EFs were
obtained for the combustion of pellets with lower ash content,
whilst the highest were registered for agro fuels. The fuel
composition, namely its ash content, was found to be a parameter with great inuence on particulate emissions. Furthermore,
the higher alkali metal content in agricultural biomass fuels in
comparison with woody pellets also contributed to the higher
PM10 EFs. Pellets made up of wood and wood wastes were
characterised by high K and ash content. The particulate emissions were found to be correlated with the gaseous emissions,
i.e., with the products of incomplete combustion. Combustion of
different fuels generated differences in the OC and EC contents of
smoke particles. Levoglucosan was the most abundant anhydrosugar in particulate matter emissions. Mannosan and galactosan were absent from almost all samples, indicating that high
temperature biomass combustion contributes little to anhydrosugar formation. The dominant inorganic species in PM10
samples showed wide uctuations depending on the fuel burned.
Potassium, usually pointed out as a biomass combustion tracer,
was not detected in some samples. The use of treated wood to
manufacture pellets was found to be problematic due to high
emissions of heavy metals, such as Zn, Pb and As.
The comparison of emissions from this study with data from the
literature showed striking differences, either in particulate emission levels or in chemical composition, between old-type appliances and modern stoves or boilers. Fuel composition and
characteristics also play a major role in emissions. The detailed
characterisation of smoke particles from appliances with high efciency is wanted since emissions can vary remarkably.
On the legislative level, national and European regulations with
more stringent particle emission limits for solid fuel heating appliances must be adopted. European-wide regulations comprising
general performance requirements to encourage further technological innovation are required. Measures leading to a strict quality
control of biofuels should be envisaged.

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

25

Table 6
Weight percentage (wt.%) of several elements in PM10.

Li
Ti
V
Cr
Mn
Cu
Zn
As
Rb
Sr
Zr
Nb
Mo
Sn
Sb
Ba
La
Ce
Nd
Sm
Bi
Pb
Th
Al
Ca
Fe
K
S
Mg
Na
P
Ni
Ga
Ge
Y
Cs
Gd
Dy
Er
Hf
W
U
B
Cd
TI

Pellets type I

Pellets type II

Pellets type III

Pellets type IV

Olive pit

Shell of pine nuts

Almond shell

0.002
0.015
0.001
0.034
0.063
0.031
0.810
0.001
0.048
0.009
0.040
0.002
0.034
0.003
0.002
0.011
<ld
<ld
<ld
<ld
<ld
0.169
<ld
0.132
1.34
0.128
17.3
4.16
0.368
1.41
0.072
0.006
<ld
<ld
<ld
0.002
<ld
<ld
<ld
0.002
0.003
<ld
0.001
0.003
<ld

0.002
0.008
0.001
0.012
0.005
0.050
5.98
0.079
0.039
0.002
0.036
0.001
0.014
0.047
0.046
0.007
<ld
<ld
<ld
<ld
0.006
5.29
<ld
0.090
0.143
0.054
12.2
1.03
0.017
4.75
0.098
0.001
0.001
<ld
<ld
0.002
<ld
<ld
<ld
0.001
0.002
<ld
0.010
0.010
0.001

0.001
0.007
<ld
0.003
0.002
0.032
3.46
0.030
0.018
<ld
0.016
0.001
0.006
0.017
0.017
0.004
<ld
<ld
<ld
<ld
0.003
2.91
<ld
0.016
<ld
1.697
<ld
1.72
5.61
0.092
6.69
<ld
<ld
<ld
<ld
0.001
<ld
<ld
<ld
0.001
0.001
<ld
0.004
0.005
<ld

0.001
0.044
0.002
0.054
0.025
0.128
3.26
0.022
0.075
0.005
0.010
0.001
0.493
0.029
0.028
0.026
0.001
0.002
0.001
0.001
0.015
1.70
0.001
0.007
<ld
5.096
<ld
1.96
3.78
0.115
5.29
0.100
0.002
0.002
0.007
0.003
0.001
0.001
0.001
<ld
0.003
0.003
0.088
0.006
0.001

<ld
0.003
<ld
0.001
0.003
0.009
0.050
0.001
0.018
0.002
0.001
<ld
0.001
<ld
<ld
0.001
<ld
<ld
<ld
<ld
<ld
0.027
<ld
0.072
0.175
0.012
20.6
2.27
0.037
0.373
0.065
<ld
<ld
<ld
<ld
0.001
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld

<ld
0.002
0.001
<ld
<ld
0.001
0.134
<ld
0.010
<ld
<ld
<ld
0.016
<ld
<ld
<ld
<ld
<ld
<ld
<ld
0.001
0.005
<ld
<ld
<ld
0.248
<ld
1.22
4.07
0.099
4.82
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld
<ld

0.001
0.010
0.001
0.015
0.007
0.035
0.064
<ld
0.035
0.003
<ld
0.001
0.125
<ld
<ld
0.007
<ld
0.001
<ld
<ld
<ld
0.011
<ld
0.047
0.154
<ld
28.1
2.59
0.044
0.362
0.130
0.025
<ld
0.002
0.002
0.001
<ld
<ld
<ld
<ld
<ld
0.001
<ld
<ld
<ld

Note: Be, Co, Pr, Eu, Ho, Tm, Lu, Ta, Tb, Yb, Sc and Se were all above the limit of detection (ld).

Acknowledgements
This work was nancially supported by AIRUSE e Testing and
development of air quality mitigation measures in Southern
Europe, LIFE 11 ENV/ES/000584, and project PTDC/AAC-AMB/
116568/2010 (FCOMP-01-0124-FEDER-019346) BiomAshTech e
Ash impacts during thermo-chemical conversion of biomass, supported by the Portuguese Foundation for Science and Technology.
Fulvio Amato is beneciary of the Juan de la Cierva postdoctoral
Grant (JCI-2012-13473) from the Spanish Ministry of Economy and
Competitiveness.
References
Alves, C., Gonalves, C., Fernandes, A.P., Tarelho, L., Pio, C., 2011. Fireplace and
woodstove ne particle emissions from combustion of western Mediterranean
wood types. Atmos. Res. 10, 692e700.
Baeza-Romero, M.T., Wilson, J.M., Fitzpatrick, E.M., Jones, J.M., Williams, A., 2010.
In situ study of soot from the combustion of a biomass pyrolysis intermediate;
eugenol; and n-decane using aerosol time of ight mass spectrometry. Energy
Fuels 24, 439e445.
Blling, A.K., Pagels, J., Yttri, K.E., Barregard, L., Sallsten, G., Schwarze, P.E., Boman, C.,
2009. Health effects of residential wood smoke particles: the importance of
combustion conditions and physicochemical particle properties. Part. Fibre

Toxicol. 6, 29. http://dx.doi.org/10.1186/1743-8977-6-29.


Calvo, A.I., Tarelho, L.A.C., Alves, C.A., Duarte, M., Nunes, T., 2014. Characterization of
operating conditions of two residential wood combustion appliances. Fuel
Process. Technol. 126, 222e232.
Carvalho, L., Wopienka, E., Lundgren, J., 2008. Challenges in small-scale combustion
of agricultural biomass fuels. Int. J. Energy Clean Environ. 9, 127e142.
Carvalho, L., Wopienka, E., Pointner, C., Lundgren, J., Verma, V.K., Haslinger, W.,
Schmidl, C., 2013. Performance of a pellet boiler red with agricultural fuels.
Appl. Energy 104, 286e296.
Cerqueira, M., Gomes, L., Tarelho, L., Pio, C., 2013. Formaldehyde and acetaldehyde
emissions from residential wood combustion in Portugal. Atmos. Environ. 72,
171e176.
Cocchi, M., Nikolaisen, L., Junginger, M., Goh, C.S., Hess, R., Jacobson, J., Ovard, L.P.,
n, D., Hennig, C., Deutmeyer, M., Schouwenberg, P.P., Marchal, D., 2011.
Thra
Global wood pellet industry. Market and trade study. In: IEA Bioenergy. Dec.
2011.
Engling, G., Carrico, C.M., Kreidenweis, S.M., Collett Jr., J.L., Day, D.E., Malm, W.C.,
Lincoln, E., Min Hao, W., Iinuma, Y., Herrmann, H., 2006. Determination of
levoglucosan in biomass combustion aerosol by high-performance anion-exchange chromatography with pulsed amperometric detection. Atmos. Environ.
40, 299e311.
Fernandes, A.P., Alves, C.A., Gonalves, C., Tarelho, L., Pio, C., Schimdl, C., Bauer, H.,
2011. Emission factors from residential combustion appliances burning Portuguese biomass fuels. J. Environ. Monit. 13, 3196e3206.
Fine, P.M., Cass, G.R., Simoneit, B.R.T., 2004. Chemical characterization of ne particle emissions from the wood stove combustion of prevalent United States tree
species. Environ. Eng. Sci. 21, 705e721.
Fitzpatrick, E.M., Jones, J.M., Pourkashanian, M., Ross, A.B., Williams, A., Bartle, K.D.,

26

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27

2008. Mechanistic aspects of soot formation from the combustion of pine wood.
Energy Fuels 22, 3771e3778.
Garcia-Maraver, A., Zamorano, M., Fernandes, U., Rabaal, M., Costa, M., 2014.
Relationship between fuel quality and gaseous and particulate matter emissions
in a domestic pellet-red boiler. Fuel 119, 141e152.
, J.,
Goh, C.S., Junginger, M., Cocchi, M., Marchal, D., Thr
an, D., Hennig, C., Heinimo
Nikolaisen, L., Schouwenberg, P.-P., Bradley, D., Hess, R., Jacobson, J., Ovard, L.,
Deutmeyer, M., 2012. Wood pellet market and trade: a global perspective.
Biofuels Bioprod. Biorening 7, 24e42.
Gonalves, C., Alves, C., Evtyugina, M., Mirante, F., Pio, C., Caseiro, A., Schmidl, C.,
Bauer, H., Carvalho, F., 2010. Characterisation of PM10 emissions from woodstove combustion of common woods grown in Portugal. Atmos. Environ. 44,
4474e4480.
Gonalves, C., Alves, C., Fernandes, A.P., Monteiro, C., Tarelho, L., Evtyugina, M.,
Pio, C., 2011. Organic compounds in PM2.5 emitted from replace and woodstove combustion of typical Portuguese wood species. Atmos. Environ. 45,
4533e4545.
Gonalves, C., Alves, C., Nunes, T., Rocha, S., Cardoso, J., Cerqueira, M., Cerqueira, M.,
, K., 2014. Organic characterisation of
Pio, C., Almeida, S.M., Hillamo, R., Teinila
PM10 in Cape Verde under Saharan dust inuxes. Atmos. Environ. 89, 425e432.
lez, J.F., Roma
na, S., Encinarb, J.M., Martnez, G., 2009. Pyrolysis of various
Gonza
biomass residues and char utilization for the production of activated carbons.
J. Anal. Appl. Pyrolysis 85, 134e141.
kiHappo, M.S., Uski, O., Jalava, P.I., Kelz, J., Brunner, T., Hakulinen, P., Ma
Paakkanen, J., Kosma, V.-M., Jokiniemi, J., Obernberger, I., Hirvonen, M.-R., 2013.
Pulmonary inammation and tissue damage in the mouse lung after exposure
to PM samples from biomass heating appliances of old and modern technologies. Sci. Total Environ. 443, 256e266.
m-Lunde
n, E., 2006.
Hedberg, E., Johansson, C., Johansson, L., Swietlicki, E., Brorstro
Is levoglucosan a suitable quantitative tracer for wood burning? comparison
with receptor modeling on trace elements in Lycksele, Sweden. J. Air Waste
Manag. Assoc. 56, 1669e1678.
Heschel, W., Rweyemamu, L., Scheibner, T., Meyer, B., 1999. Abatement of emissions
in small-scale combustors through utilisation of blended pellet fuels. Fuel
Process. Technol. 61, 223e242.
Ingwald, O., Gerold, T., 2006. Recent developments concerning pellet combustion
technologies e a review of Austrian developments. In: Swedish Bioenergy Association (Ed.), Proceedings of the 2nd World Conference on Pellets, pp. 31e40.
nko
ping, Sweden.
Jo
Johansson, L.S., Leckner, B., Gustavsson, L., Cooper, D., Tullin, C., Potter, A., 2004.
Emission characteristics of modern and old-type residential boilers red with
wood logs and wood pellets. Atmos. Environ. 38, 4183e4195.
Kuo, L.-J., Herbert, B.E., Louchouarn, P., 2008. Can levoglucosan be used to characterize and quantify char/charcoal black carbon in environmental media? Org.
Geochem. 39, 1466e1478.
Kuo, L.-J., Louchouarn, P., Herbert, B.E., 2011. Inuence of combustion conditions on
yields of solvent-extractable anhydrosugars and lignin phenols in chars: implications for characterizations of biomass combustion residues. Chemosphere
85, 797e805.
Lamberg, H., Nuutinen, K., Tissari, J., Ruusunen, J., Yli-Piril
a, P., Sippula, O.,
, K., Saarnio, K., Hillamo, R.,
Tapanainen, M., Jalava, P., Makkonen, U., Teinila
Hirvonen, M.-R., Jokiniemi, J., 2011a. Physicochemical characterization of ne
particles from small-scale wood combustion. Atmos. Environ. 45, 7635e7643.
Lamberg, H., Sippula, O., Tissari, J., Jokiniemi, J., 2011b. Effects of air staging and load
on ne-particle and gaseous emissions from a small-scale pellet boiler. Energy
Fuels 25, 4952e4960.
n, a., Torvela, T., Kaivosoja, T., Tapanainen, M.,
Leskinen, J., Tissari, J., Uski, O., Vire
, K., Saarnio, K., Hillamo, R., Hirvonen, M.-R.,
Jalava, P., Makkonen, U., Teinila
Jokiniemi, J., 2014. Fine particle emissions in three different combustion conditions of a wood chip-red appliance e particulate physico-chemical properties and induced cell death. Atmos. Environ. 86, 129e139.
ndez, M.J., Esteban, L.S., 2009. Optimization of pelletisation and
Mediavilla, I., Ferna
combustion in a boiler of 17.5 kWth for vine shoots and industrial cork residue.
Fuel Process. Technol. 90, 621e628.
n, S., Rojas, C.V., Nogales, S., 2012. CharMiranda, T., Arranz, J.I., Montero, I., Roma
acterization and combustion of olive pomace and forest residue pellets. Fuel
Process. Technol. 103, 91e96.
Monteiro, E., Mantha, V., Rouboa, A., 2012. Portuguese pellets market: analysis of
the production and utilization constrains. Energy Policy 42, 129e135.
Obaidullah, M., Bram, S., Verma, V.K., Ruyck, J. De, 2012. A review on particle
emissions from small scale biomass combustion. Int. J. Renew. Energy Res. 2 (1).
Obernberger, I., Thek, G., 2004. Physical characterisation and chemical composition
of densied biomass fuels with regard to their combustion behavior. Biomass
Bioenergy 27, 653e669.

Ohman, M., Nordin, A., Hedman, H., Jirjis, R., 2004. Reasons for slagging during
stemwood pellet combustion and some measures for prevention. Biomass
Bioenergy 27, 597e605.
llstrand, J., 2006. Low emissions from wood burning in an ecolabelled
Olsson, M., Kja
residential boiler. Atmos. Environ. 40, 1148e1158.
Orasche, J., Seidel, T., Hartmann, H., Schnelle-kreis, J., Chow, J.C., Ruppert, H.,
Zimmermann, R., 2012. Comparison of emissions from wood combustion. Part
1: emission factors and characteristics from different small-scale residential
heating appliances considering particulate ne particle and gaseous emissions
from normal and smouldering wood combustion in a conventional masonry
heater matter and polycyclic aromatic hydrocarbon (PAH)-related toxicological

potential. Energy Fuels 26, 6695e6704.


Ozgen, S., Caserini, S., Galante, S., Giugliano, M., Angelino, E., Marongiu, A.,
Hugony, F., Morreale, C., 2014. Emission factors from small scale appliances
burning wood and pellets. Atmos. Environ. 94, 144e153.

Pettersson, E., Lindmark, F., Ohman,


M., Nordin, A., Westerholm, R., Boman, C., 2010.
Design changes in a xed-bed pellet combustion device: effects of temperature
and residence time on emission performance. Energy Fuels 24, 1333e1340.
m, D., Nordin, A., 2011. Stove perPettersson, E., Boman, C., Westerholm, R., Bostro
formance and emission characteristics in residential wood log and pellet
combustion, part 2: wood stove. Energy Fuels 25, 315e323.
schl, U., 2005. Atmospheric aerosols: composition, transformation, climate and
Po
health effects. Angew. Chem. Int. Ed. Engl. 44, 7520e7540.
Querol, X., Alastuey, A., Rodriguez, S., Plana, F., Ruiz, C.R., Cots, N., 2001. PM10 and
PM2.5 source apportionment in the Barcelona Metropolitan area, Catalonia,
Spain. Atmos. Environ. 35, 6407e6419.
Riva, G., Pedretti, E.F., Toscano, G., Duca, D., Pizzi, A., 2011. Determination of polycyclic aromatic hydrocarbons in domestic pellet stove emissions. Biomass
Bioenergy 35, 4261e4267.
Rose, D., Nowak, A., Achtert, P., Wiedensohler, A., Hu, M., Shao, M., Zhang, Y.,
schl, U., 2010. Cloud condensation nuclei in polluted air and
Andreae, M.O., Po
biomass burning smoke near the mega-city Guangzhou, Chinadpart 1: sizeresolved measurements and implications for the modeling of aerosol particle
hygroscopicity and CCN activity. Atmos. Chem. Phys. 10, 3365e3383.
Sander, B., 1997. Properties of Danish biofuels and the requirements for power
production. Biomass Bioenergy 12, 177e183.
Schmidl, C., Luisser, M., Padouvas, E., Lasselsberger, L., Rzaca, M., Ramirez-Santa
Cruz, C., Handler, M., Peng, G., Bauer, H., Puxbaum, H., 2011. Particulate and
gaseous emissions from manually and automatically red small scale combustion systems. Atmos. Environ. 45, 7443e7454.
, P., Berner, A., Bauer, H., Kasper-Giebl, A.,
Schmidl, C., Marr, I.L., Caseiro, A., Kotianova
Puxbaum, H., 2008. Chemical characterisation of ne particle emissions from
wood stove combustion of common woods growing in mid-European Alpine
regions. Atmos. Environ. 42, 126e141.
Sippula, O., Hyto, K., Tissari, J., Raunemaa, T., Jokiniemi, J., 2007. Effect of wood fuel
on the emissions from a top-feed pellet stove. Energy Fuels 21, 1151e1160.
Skotland, C.H., 2009. Measurement of Temperature Conditions in Grate Zone of a 1
MW Wood-pellets Boiler Fired with High Ash Content Wood-pellets (Master of
Science in Energy and Environment). Norwegian University of Science and
Technology.
Tang, X., Bai, Y., Duong, A., Smith, M.T., Li, L., Zhang, L., 2009. Formaldehyde in
China: production, consumption, exposure levels, and health effects. Environ.
Int. 35, 1210e1224.
Telmo, C., Lousada, J., Moreira, N., 2010. Proximate analysis, backwards stepwise
regression between gross caloric value, ultimate and chemical analysis of
wood. Bioresour. Technol. 101, 3808e3815.
nena, K., Lyyra
nen, J., Jokiniemi, J., 2007. A novel eld measurement
Tissari, J., Hyto
method for determining ne particle and gas emissions from residential wood
combustion. Atmos. Environ. 41, 8330e8344.
nen, J., Hyto
nen, K., Sippula, O., Tapper, U., Frey, A., Saarnio, K.,
Tissari, J., Lyyra
Pennanen, A.S., Hillamo, R., Salonen, R.O., Hirvonen, M.-R., Jokiniemi, J.,
2008a. Fine particle and gaseous emissions from normal and smouldering
wood combustion in a conventional masonry heater. Atmos. Environ. 42,
7862e7873.
Tissari, J., Sippula, O., Kouki, J., Vuorio, K., Jokiniemi, J., 2008b. Fine particle and gas
emissions from the combustion of agricultural fuels red in a 20 kW burner.
Energy Fuels 22, 2033e2042.
Tissari, J., 2008. Fine Particle Emissions from Residential Wood Combustion
(Doctoral thesis). Department of Environmental Science. University of Kuopio.
Toscano, G., Duca, D., Amato, A., Pizzi, A., 2014. Emission from realistic utilization of
wood pellet stove. Energy 68, 644e650.
n, A., La
hde, A.,
Torvela, T., Tissari, J., Sippula, O., Kaivosoja, T., Leskinen, J., Vire
Jokiniemi, J., 2014. Effect of wood combustion conditions on the morphology of
freshly emitted ne particles. Atmos. Environ. 87, 65e76.
Turpin, B.J., Lim, H.-J., 2001. Species contributions to PM2.5 mass concentrations:
revisiting common assumptions for estimating organic mass. Aerosol Sci.
Technol. 35, 602e610.
kiUski, O., Jalava, P.I., Happo, M.S., Leskinen, J., Sippula, O., Tissari, J., Ma
Paakkanen, J., Jokiniemi, J., Hirvonen, M.-R., 2014. Different toxic mechanisms
are activated by emission PM depending on combustion efciency. Atmos.
Environ. 89, 623e632.
Verma, V.K., Bram, S., Gauthier, G., De Ruyck, J., 2011. Evaluation of the performance
of a multi-fuel domestic boiler with respect to the existing European standard
and quality labels: part-1. Biomass Bioenergy 35, 80e89.
Verma, V.K., Bram, S., Delattin, F., Laha, P., Vandendael, I., Hubin, A., De Ruyck, J.,
2012. Agro-pellets for domestic heating boilers: standard laboratory and real
life performance. Appl. Energy 90, 17e23.
Vicente, A., Alves, C., Monteiro, C., Nunes, T., Mirante, F., Cerqueira, M., Calvo, A.,
Pio, C., 2012. Organic speciation of aerosols from wildres in central Portugal
during summer 2009. Atmos. Environ. 57, 186e196.
Vicente, E.D., Vicente, A.M., Bandowe, B.A.M., Alves, C.A., 2015. Particulate phase
emission of parent polycyclic aromatic hydrocarbons (PAHs) and their derivatives (alkyl-PAHs, oxygenated-PAHs, azaarenes and nitrated-PAHs) from
manually and automatically red combustion appliances. Air Qual. Atmos.
Health. http://dx.doi.org/10.1007/s11869-015-0364-1 (in press).
Wasson, S.J., Linak, W.P., Gullett, B.K., King, C.J., Touati, A., Huggins, F.E., Chen, Y.,

E.D. Vicente et al. / Atmospheric Environment 120 (2015) 15e27


Shah, N., Huffman, G.P., 2005. Emissions of chromium, copper, arsenic, and
PCDDs/Fs from open burning of CCA treated wood. Environ. Sci. Technol. 39,
8865e8876.
Wiinikka, H., 2015. High Temperature Aerosol Formation and Emission Minimisation during Combustion of Wood Pellets (Doctoral thesis).
Yokelson, R.J., Susott, R., Ward, D.E., Reardon, J., Grifth, D.W.T., 1997. Emissions
from smoldering combustion of biomass measured by open-path Fourier
transform infrared spectroscopy. J. Geophys. Res. 102 (D15), 18865e18877.

27

Zhang, L., Steinmaus, C., Eastmond, D.A., Xin, X.K., Smith, M.T., 2009. Formaldehyde
exposure and leukemia: a new meta-analysis and potential mechanisms. Mutat.
Res. 681, 150e168.
Zhang, X., Li, J., Yang, W., Blasiak, W., 2011. Formation mechanism of levoglucosan
and formaldehyde during cellulose pyrolysis. Energy Fuels 25, 3739e3746.
Zhang, Y., Obrist, D., Zielinska, B., Gertler, A., 2013. Particulate emissions from
different types of biomass burning. Atmos. Environ. 72, 27e35.

S-ar putea să vă placă și