You are on page 1of 13

Science of the Total Environment 612 (2018) 636–648

Contents lists available at ScienceDirect

Science of the Total Environment

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

Chemical composition and speciation of particulate organic matter from


modern residential small-scale wood combustion appliances
Hendryk Czech a, Toni Miersch a, Jürgen Orasche a,b,c, Gülcin Abbaszade b,c, Olli Sippula c,d, Jarkko Tissari d,
Bernhard Michalke e, Jürgen Schnelle-Kreis b,c, Thorsten Streibel a,b,c,⁎,
Jorma Jokiniemi c,d, Ralf Zimmermann a,b,c
a
Joint Mass Spectrometry Centre, Chair of Analytical Chemistry, Institute of Chemistry, University of Rostock, 18059 Rostock, Germany
b
Joint Mass Spectrometry Centre, Cooperation Group “Comprehensive Molecular Analytics” (CMA), Helmholtz Zentrum München – German Research Centre for Environmental Health, 85764
Neuherberg, Germany
c
Helmholtz Virtual Institute of Complex Molecular Systems in Environmental Health (HICE), Germany
d
Fine Particle and Aerosol Technology Laboratory, Department of Environmental and Biological Sciences, University of Eastern Finland, FIN-70211 Kuopio, Finland
e
Research Unit Analytical BioGeochemistry, Helmholtz Zentrum München – German Research Centre for Environmental Health, 85764 Neuherberg, Germany

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• PM2.5 composition from two modern


wood combustion appliances
• Advanced targeted and non-targeted
mass spectrometric techniques
• Emission factors for N 100 organic parti-
cle constituents
• Low abundance of established wood
combustion markers
• Slow ignition can shift emission pattern
compared to regular combustion.

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

Article history: Combustion technologies of small-scale wood combustion appliances are continuously developed decrease emis-
Received 27 June 2017 sions of various pollutants and increase energy conversion. One strategy to reduce emissions is the implementa-
Received in revised form 23 August 2017 tion of air staging technology in secondary air supply, which became an established technique for modern wood
Accepted 26 August 2017
combustion appliances. On that account, emissions from a modern masonry heater fuelled with three types of
Available online 1 September 2017
common logwood (beech, birch and spruce) and a modern pellet boiler fuelled with commercial softwood pellets
Editor: D. Barcelo were investigated, which refer to representative combustion appliances in northern Europe In particular, empha-
sis was put on the organic constituents of PM2.5, including polycyclic aromatic hydrocarbons (PAHs), oxygenated
Keywords: PAHs (OPAHs) and phenolic species, by targeted and non-targeted mass spectrometric analysis techniques. Com-
Diagnostic ratios pared to conventional wood stoves and pellet boilers, organic emissions from the modern appliances were re-
Wood smoke duced by at least one order of magnitude, but to a different extent for single species. Hence, characteristic
PAH ratios of emission constituents and emission profiles for wood combustion identification and speciation do not
OPAH hold for this type of advanced combustion technology. Additionally, an overall substantial reduction of typical
Phenolics
wood combustion markers, such as phenolic species and anhydrous sugars, were observed. Finally, it was
Photoionisation mass spectrometry
found that slow ignition of log woods changes the distribution of characteristic resin acids and phytosterols as

⁎ Corresponding author at: Joint Mass Spectrometry Centre, Chair of Analytical Chemistry, University of Rostock, 18059 Rostock, Germany.

http://dx.doi.org/10.1016/j.scitotenv.2017.08.263
0048-9697/© 2017 Elsevier B.V. All rights reserved.
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 637

well as their thermal alteration products, which are used as markers for specific wood types. Our results should
be considered for wood combustion identification in positive matrix factorisation or chemical mass balance in
northern Europe.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction of future emission scenarios to estimate benefits of emission abatement


technologies on air quality (Fountoukis et al., 2014).
In many countries of Europe as well as in North America, the resi-
dential combustion of biomass, in particular wood, as a renewable ener- 2. Material and methods
gy source is encouraged by legislation to decrease the dependence on
fossil fuels and contribution to global warming. Conventional logwood 2.1. Experimental setup
stoves, pellet stoves and pellet boilers have been recognised to be sub-
stantial emitters of greenhouse gases, black or elemental carbon and 2.1.1. Combustion appliances
brown carbon (Evtyugina et al., 2014; Martinsson et al., 2015; Orasche The emissions of two small-scale wood combustion appliances with
et al., 2012), which counteract the widely considered CO2-neutrality advanced secondary air supply for residential heating were investigat-
and thus affect the climate (Andreae and Ramanathan, 2013). Further- ed. The modern masonry heater (Hiisi 4, Tulikivi Ltd., Finland) com-
more, residential wood burning causes elevated ambient concentra- prises of a massive soap stone of app. 1.3 tons for slow heat release, an
tions of inhalable particulate matter (PM) with an aerodynamic upright enclosed firebox and a double glass window door. In the firebox,
diameter below 2.5 μm (PM2.5) (Fuller et al., 2013), which has been as- the secondary air is supplied through panels with small rifts in the
sociated with toxicological responses, such as genotoxicity, cytotoxicity, upper combustion chamber (air staging) to generate a secondary com-
oxidative stress, systemic inflammation and cardiovascular diseases bustion zone, which was shown to substantially reduce emissions of CO,
(Croft et al., 2017; Miljevic et al., 2010; Naeher et al., 2007; Sehlstedt VOCs and organic matter (Nuutinen et al., 2014). The modern masonry
et al., 2010). Also, the emissions of volatile organic compounds (VOCs) heater was operated at approximate nominal load.
from wood combustion cover a substantial potential for secondary or- The second appliance is an automatically-fired top-feed pellet boiler
ganic aerosol (SOA) formation (Bruns et al., 2016), which may have dif- (PZ25RL, Biotech Energietechnik GmbH, Austria), which was continu-
ferent health effects than the primary emissions (Künzi et al., 2013; ously operated at its nominal power of 25 kW. A more detailed descrip-
Nordin et al., 2015) and additionally contributes to ambient PM2.5 tion of the pellet boiler and the effect of air staging can be found in
levels. Lamberg et al. (2011b).
In order to tackle these emissions, new and more complete combus- In the following, the term “modern” refers to wood combustion ap-
tion technologies were developed which may significantly reduce the pliances with air staging and “conventional” without.
toxicity potential of primary wood combustion aerosols (Jalava et al., This study was part of experiments by Helmholtz Virtual Institute of
2012). Secondary air supply through air staging became an established Complex Molecular System in Environmental Health (HICE), which
emission abatement strategy, which substantial reduce the emissions of aims to explore biological effects of emissions on human lung epithelial
several organic and inorganic pollutants (Nuutinen et al., 2014). Air cells (Kanashova et al., 2017). In particular, emissions from state-of-the-
staging is characterised by splitting the total combustion air into art combustion appliances and emerging fuels are of interest as they
under-stoichiometric air-to-fuel ratios for primary air (λ b 1) and may represent future scenarios. The two combustion appliances were
consecutive addition of low excess secondary air (λ N 1.5), which is chosen due to their known low emissions of bulk components from pre-
well-mixed with the pyrolysis gases through different feedings. Thus, vious studies (Nuutinen et al., 2014; Lamberg et al., 2011b) and their
high temperatures and more complete combustion are achieved representativeness for single houses in terms of energy output and mar-
(Nussbaumer, 2003). However, some pollutants correlate inversely ket availability.
with increasing secondary air flow rates, such as the trade-off between
CO and NOx (Khodaei et al., 2017). 2.1.2. Fuels and combustion procedure
Emissions from a masonry heater and a pellet boiler, equipped with The modern masonry heater was fuelled by beech (Fagus sylvatica)
air staging technology, were previously investigated regarding PM, CO, and spruce (Picea abies), which are common firewood in central
NOx (NO + NO2), total VOC (Lamberg et al., 2011a; Lamberg et al., Europe, as well as birch (Betula pubescens), which is a typical firewood
2011b; Nuutinen et al., 2014) and VOC composition (Czech et al., in northern Europe. In Table 1, physico-chemical properties of the
2017; Czech et al., 2016; Reda et al., 2015), as well as their SOA forma- three types of firewood are summarised. In total 15 kg of logwood
tion potential (Kari et al., 2017; Tiitta et al., 2016). In this study, an addi- was burned in a single experiment, split up in six consecutive batches
tional detailed characterisation of the PM2.5 composition is presented of 2.5 kg each. The ignition of the first batch in the cold stove was carried
with emphasis on the organic constituents, whose emission factors out from top down with 150 g of small wood sticks/chips. After 35 min,
(EFs) are compared to conventional and other modern wood stoves. the next batch was put in the wood combustion residues for self-
The implications for source apportionment studies of the results on ignition and burned for further 35 min. Subsequent to the sixth batch,
chemical patterns of emissions are also discussed, especially the impli- remaining ember was stoked and the secondary air supply was blocked
cations for selecting common diagnostic ratios to trace sources and according to manufacture instructions for 30 min to cover a total exper-
quantify source contributions from biomass burning to ambient PM imental time of 4 h.
concentrations. Moreover, differences of EFs and emission profiles be- In two (spruce* and birch*) of the total log wood combustion exper-
tween wood types are related to effects observed in toxicological stud- iments, a slow ignition of the first batch was observed. Due to qualita-
ies (Kasurinen et al., 2017). tive and quantitative differences in emissions of these experiments
Although the masonry heater and the pellet boiler of this study are compared to the other ones, the effect of slow ignition on the total EFs
only representative domestic combustion appliances in northern is discussed more in detail in Section 3.5.
Europe, it was pointed out that especially features of regional emission The pellet boiler was operated with commercial softwood pellets
activity are important for emission inventories and source apportion- (produced from pine and spruce stem wood, physico-chemical properties
ment (Hellén et al., 2008; Pastorello et al., 2011). In the end, the EFs in Table 1) under stable conditions over 4 h and connected to a heat ex-
from the two combustion appliances can be also involved in simulations changer. Those boilers are usually connected to a reservoir of water for
638 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

Table 1
Physico-chemical properties of the wood fuels.

Parameter Unit Limit of quantification Method Beech logs Birch logs Spruce logs Softwood pellets

Moisture % (w/w) 0.1 DIN EN 14774-2 9.0 7.2 7.4 7.3


Ash content % (w/w, dry) 0.1 DIN EN 14775 1.3 0.69 0.58 0.36
C % (w/w, dry) 0.2 DIN EN 15104 50.3 51.0 52.0 51.7
H % (w/w, dry) 0.1 DIN EN 15104 5.8 6.0 5.9 6.0
N % (w/w, dry) 0.05 DIN EN 15104 0.36 0.40 0.36 0.29
O % (w/w, dry) Calculated 42.3 41.9 41.1 41.6
S % (w/w, dry) 0.005 DIN EN 15289 0.037 0.006 0.009 0.006
Cl % (w/w, dry) 0.005 DIN EN 15289 b0.005 b0.005 0.005 0.006
Zn mg/kg (dry) 1 DIN EN ISO 17294-2 5 33 25 14
Si mg/kg (dry) 1 DIN EN ISO 11885 6.3 17.5 5.8 14.5
K mg/kg (dry) 1 DIN EN ISO 11885 14.5 9.3 14.5 12.1
Net heating value kJ/kg 200 DIN EN 14918 17,790 18,140 18,640 18,380

warm water supply and therefore have to constantly produce heat in con- ImproveA protocol (Chow et al., 2007). During the four thermal sub-
trast to pellet stoves, which refers to room heating appliances. A more de- fractions of OC (OC1 to OC4) with upper temperature limits of 140 °C,
tailed description of the combustion experiments can be found in Reda 280 °C, 480 °C and 580 °C, approximately 8% of the total flow enters
et al. (2015). the TOFMS through a deactivated transfer capillary (inner diameter of
320 μm), which is connected to the carbon analyser oven by a modified
2.2. Instrumentation quartz cross stepwise heated from 230 °C to 245 °C to prevent conden-
sation (Grabowsky et al., 2011). In the ion source, UV radiation of
2.2.1. Gas measurements 266 nm which is provided by the fourth harmonic generation of an
Flue gases were led from the firebox to the upper combustion cham- Nd:YAG laser (Spitlight400, Innolas, Germany; 20 Hz repetition rate,
ber and then downwards through side ducts into the stack. The stack 1064 nm fundamental radiation, energy of 1 mJ at 266 nm) ionises se-
was placed below a hood, and draught was regulated with two lectively desorbed aromatic constituents of the particles in the REMPI
dampers. The target value for the pressure in the stack was (12.0 ± process. In addition to the optical selectivity, REMPI denotes a soft
0.5) Pa below ambient pressure. CO, CO2, NOx and O2 were continuously ionisation technique leading to predominantly molecular ions and low
measured by a gas analyser system (ABB, Limas 11 UV, Switzerland). yields of fragments (Boesl, 2000). The generated ions were subsequent-
Organic gaseous carbon (OGC), a quantity for total organic vapours, ly analysed by a TOFMS (compact reflectron time-of-flight spectrometer II,
were quantified by a flame ionisation detector (ABB, Multi-FID 14, Stefan Kaesdorf Geräte für Forschung und Industrie, Germany) with a
Switzerland), which was calibrated against propane. All gaseous emis- mass resolution of 1000 at m/z 78 and 1 s time resolution. Finally, the
sions were measured directly from undiluted stack gas through an insu- obtained mass spectra were summed according to their occurrence in
lated and externally heated sampling line at 180 °C. the four fractions of OC.
From the analysis of O2 in the flue gas, emission factors (EFs) were For targeted analysis, in-situ derivatisation direct thermal desorp-
calculated based on the instruction of the Finnish Standard Association tion gas chromatography time-of-flight mass spectrometry (IDTD-GC-
method SFS 5624 as described in Reda et al. (2015). TOFMS) with electron ionisation (Orasche et al., 2011) was applied to
quantify in total 105 particle-bound organic compounds covering the
2.2.2. Filter sampling of PM2.5 substance classes of polycyclic aromatic hydrocarbons (PAHs), oxygen-
The emitted aerosol of the wood combustion appliances was ated PAHs (OPAHs), hydroxy-PAH (OH-PAHs), polycyclic aromatic
isokinetically sampled from the stack, diluted at a dilution ratio of 40 sulphur-containing hydrocarbons (PASH), anhydrous sugars, phenolic
by using a porous-tube/ejector dilutor (Venacontra, Finland) and cooled species, resin acids and phytosterols. The derivatisation is based on
to room temperature to enable condensation of semi-volatile species. silylation with N-Methyl-N-trimethylsilyltrifluoroacetamide (MSTFA)
Subsequently, particles were segregated to an aerodynamic diameter during the step of thermal desorption from quartz fibre filter. Detected
smaller than 2.5 μm and isokinetically sampled on quartz fibre filters compounds were identified by library match of electron ionisation spec-
(T293, Munktell, Sweden) over the total experiment duration of 4 h. Fi- tra as well as retention index and quantified by isotope-labelled internal
nally, the filter samples were stored at −20 °C until analysis. An over- standards of the same substance or chemically similar substances.
view about the total experimental setup can be found in Weggler et al.
(2016). Please note that gas-particle partitioning, which affects the 2.2.4. Inorganic PM2.5 constituents
emission factors, are dependent on the applied dilution technique and Samples of particulate matter were reconditioned prior to elemental
dilution ratio (Nuutinen et al., 2014). The applied dilution ratio around analysis by inductively-coupled plasma optical emission spectrometry
40 has been widely used for sampling PM emissions, but it was demon- (ICP-OES, Spectro Ciros Vision, SPECTRO Analytical Instruments GmbH
strated that towards higher dilutions the sampled PM mass decreases & Co., Germany) as described in Sections 1.1 to 1.3 of the Supplementary
substantially (Lipsky and Robinson, 2006). Thus, the presented EFs can data. In total 19 elements, including Al, B, Ba, Bi, Ca, Cd, Cr, Cu, Fe, K, Li,
be regarded as an upper limit. Mn, Mo, Na, Pb, S, Ti, W and Zn were quantified and presented if at
least 50% of the samples of one wood type contained concentrations
2.2.3. Organic PM2.5 constituents higher than the limit of quantification.
The analysis of the organic fraction of the PM2.5 emissions was
carried out by targeted and untargeted mass spectrometric techniques. 2.2.5. Data analysis and statistics
Thermal/optical carbon analysis (TOCA) coupled to resonance- For all statistical analysis, including cluster analysis and analysis of
enhanced multi-photon ionisation time-of-flight mass spectrometry variance (ANOVA), Matlab® basic functions and Statistic Toolbox
(REMPI-TOFMS) refers to an untargeted technique and enables to quan- (R2014b, The MathWorks Inc., Massachusetts, USA) was used. In case
tify organic (OC) and elemental carbon (EC), but also to investigate the of significance at a level of 0.05, ANOVA was followed by Bonferroni cor-
molecular composition of the thermal sub-fractions related to OC (Diab rection method to avoid accumulation of type I error at multiple testing.
et al., 2015). A filter punch of 0.5 cm2 was placed into thermal-optical If not otherwise indicated, reported p-values in the following refer to re-
carbon analyser (DRI model 2001a) and analysed following the sults of multiple comparison tests. A result is called significant if p-
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 639

values appear below the significance level α of 0.05. A summary of fold


changes and p-values for all comparisons between the four wood types
is depicted in volcano diagrams (Figs. S1–S6). Compounds with concen-
trations below the limit of quantification in each single experiment for
one wood fuel were not considered in the visualisation.

3. Results & discussion

3.1. Gaseous emissions

Regarding the main gaseous constituents in the flue gas of the ma-
sonry heater, only marginal differences between the mean EFs of the
different wood types could be observed, which is also reflected by the
similar modified combustion efficiencies (MCE = ΔCO2 / (ΔCO2 +
ΔCO; Δ denotes background correction) in the range of 0.983 (spruce)
to 0.988 (birch). In contrast to that, the more controlled and continuous
combustion in the pellet boiler resulted in about one order of magnitude
lower emissions of CO and OGC, despite lower excess flue gas O2 con-
centration. Generally, emissions of NOx predominantly result from
fuel nitrogen and appear in the range of 83 to 109 mg MJ−1, which is
comparable to other wood stoves and pellet boilers of older types
(Orasche et al., 2012). A comparison of the gaseous and VOC emissions
from these two wood combustion appliances with literature data from
logwood stoves, pellet stoves and boiler of older types was already de-
scribed by Czech et al. (2016) and Czech et al. (2017).

3.2. Inorganic PM2.5 constituents


Fig. 1. Scatter plot of the Zn (upper panel) and K (lower panel) content of the wood fuels
Inorganic emissions from the wood combustion appliances are af- beech (circles), birch (diamonds), spruce (triangles) and pellets (pentagrams) versus EF of
Zn and K. Numbers in the lower panel refers to sulphur-to-chlorine ratio of the wood fuels.
fected by the content of the respective element in the fuel, the fuel
bed temperature, the volatility of the element and concentrations of
possible binding partners. K, S, Na and Zn denote the most abundant in- volatile KCl, whereas higher fuel-sulphur favours the formation of less
organic particle constituents. In particular, the highest emissions of al- volatile K2(SO4) (Sippula et al., 2007; Sippula et al., 2008). In agreement
kali, alkaline earth and transient metals were detected in emissions with that, a significant moderate correlation (r = 0.55, p = 0.02) was
from birch wood and pellet combustion. In all multiple comparisons found for the ratio of Cl to S and K. However, for some fuels only
after significant ANOVA results, significant elevated EFs for Zn, Cd and upper limits for the ratio of Cl to S could be used because of Cl concen-
W were found for birch wood with a fold change above 1.5. The pellet trations below the limit of quantification (Table 1). Furthermore, the
boiler operated on softwood pellets revealed significantly increased Si content of the fuel can decrease the release of K through formation
EFs for Na, K and S (Figs. S3, S5 and S6). Interestingly, Bi, Li and Mo of low-volatile potassium silicates, which remain in the bottom ash
were even only detected in particles from pellet combustion. On the (Sommersacher et al., 2012). However, a correlation between the ratio
contrary, Cd was only detected in the particles from logwood combus- of Si to K of the fuel and K in the emissions could not be found.
tion although the Cd content of the woods was below the limit of quan-
tification of 0.2 mg/kg for all investigated wood fuels. 3.3. Targeted analysis of carbonaceous PM2.5 constituents
Of all analysed metals, only Zn showed a steady increase and signif-
icant linear correlation (r = 0.907, p = 1.2 · 10−6) between particle- 3.3.1. Organic and elemental carbon
bound Zn emissions and increasing content in the fuel, independently The combustion of all types of logwood in the masonry heater led to
from the combustion appliance (Fig. 1). Fractions of 19%, 37%, 22% and more than one order of magnitude higher emissions of both OC and EC
18% of the fuel-Zn are released in particle emissions for beech, birch, compared to the pellet boiler (OC: 0.15 mg MJ−1; EC: 0.27 mg MJ−1).
spruce and pellets, respectively, which agrees well with earlier observa- However, some significant differences were recognised between the
tions on Zn release from various wood-fired combustion appliances logwood types. The combustion of the hardwoods beech and birch
(Lamberg et al., 2011b; Sippula et al., 2017). Particulate Zn, solely pres- emitted more total carbon (TC) than spruce, but birch produced
ent as ZnO in wood smoke particles, is of great interest because it has 47 mg MJ−1 EC compared to 28 mg MJ−1 EC for beech and only
been found to be an important driver of cytotoxicity (Uski et al., 13 mg MJ−1 EC for spruce. Regarding OC, a three-fold higher EF was
2015). For all possible comparisons, birch wood led always to signifi- found for beech (20 mg MJ−1) compare to the equal EFs for birch and
cantly enhanced Zn emissions with a minimum fold change of 4. On spruce (6.1 mg MJ−1). Generally, carbonaceous particulate emissions
that account, Zn emission can be lowered by preferring Zn-poor fire- were substantially lower compared to wood combustion appliances
wood or detaching the bark in which Zn is enriched (Wiinikka et al., without air-staging (Calvo et al., 2015; Orasche et al., 2012) and compa-
2013). rable to other advanced stoves with primary, secondary and post-
K in PM2.5 emissions, which used as one indicator for biomass com- combustion air supply (Tschamber et al., 2016).
bustion in source apportionment studies (Watson et al., 2001), did not The ratio of OC to EC (OC/EC) has been widely applied in atmospher-
correlate with fuel-K. The release of K is generally affected by the avail- ic sciences for the apportionment of emission sources, such as to bio-
ability and ratio of Si to K and S to Cl. These factors have been found to mass burning, combustion of fossil fuel or SOA. In particular, the
influence the partitioning of K between coarse-sized ashes (bottom minimum OC/EC plays an important role to estimate the amount of
ash and coarse fly ash) and fine fly ashes, which forms part of the fine SOA on ambient particles while OC/EC b 1 was associated with fossil
particulate emission (Sippula et al., 2017; Sommersacher et al., 2012). fuel combustion (Pio et al., 2011). Although it has been demonstrated
A higher content of Cl in the wood fuel leads to the formation of more that the air-staging of the investigated modern masonry heater has a
640 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

only small effect on the emission of total carbon in PM2.5 (Lamberg not work for automatically-fired combustion appliances even though
et al., 2011a; Nuutinen et al., 2014), the data reveals consistently OC/ levoglucosan and mannosan could be detected (Schmidl et al., 2011).
EC b 1 for each single combustion experiment, which might lead to mis- For the pellet boiler, LGS/MNS of 24.8 ± 8.1 exceeded the higher LGS/
interpretation of ambient air studies in areas where this type of wood MNS for hardwood although the pellets were solely comprised of
stoves are used. Moreover, the combustion of birch logs generated par- softwood.
ticles with the lowest OC/EC of 0.13 ± 0.02, which is significantly differ- Galactosan was the lowest abundant anhydrous sugar without sig-
ent compared to beech logs (0.71 ± 0.14) and pellets (0.65 ± 0.23). nificant differences between the three logwoods. Regarding the pellet
Compared to spruce (0.48 ± 0.13), the difference of OC/EC from birch boiler particles, galactosan concentrations in each single combustion ex-
logs was not significant, but substantial (p = 0.058). periment were found below the limit of quantification.
In this context it must be mentioned that the amount of carbonates
(inorganic carbon, IC), which were not analysed in this study, may bias 3.3.3. Phenolic species
the OC/EC results. In particular for the pellet boiler IC can cover the same The EFs of 23 quantified phenolic species cover a range from below
range as OC or EC (Lamberg et al., 2011a; Lamberg et al., 2011b). How- limit of quantification to 36 μg MJ−1 with molecular patterns according
ever, in TOCA the contribution of IC either to OC or EC depends on the to the current knowledge of lignin monomers in different wood species.
decomposition temperature and therefore on the cation of the Syringyl and sinapyl alcohols and their derivatives on particulate emis-
carbonate. sions are associated with hardwoods, whereas softwood burning emis-
sions contain higher relative amounts of coniferyl-derived compounds.
3.3.2. Anhydrous sugars Guaiacol as another abundant building block of the wood is emitted by
Among all anhydrous sugars, levoglucosan denotes the most rele- both wood families in equal amounts (Simoneit, 2002). These three
vant one because of its established application as marker compound basic rules also hold for the performed experiments. Concerning total
for biomass combustion (Simoneit, 2002). It originates from the thermal particle-bound phenolic species, beech wood showed the highest EF
decomposition of cellulose (Shafizadeh, 1968) and generally refers to (72 μg MJ− 1), followed by birch (20 μg MJ− 1), spruce (8.4 μg MJ−1)
one of the most abundant organic compounds in wood combustion par- and pellets (0.59 μg MJ−1). In the pellet emissions only 5 of the 23 pre-
ticles. Despite the large differences in mean EFs of levoglucosan be- sented phenolic species appeared above the limit of quantification,
tween the combustion experiments, ranging from 113 μg MJ− 1 which complicates the identification of pellet boiler emissions in atmo-
(beech) to 10 μg MJ− 1 (pellet), no significance was identified by spheric studies. Furthermore, the question of possibly enhanced toxicity
ANOVA. The highest concentrations of the levoglucosan are emitted of the released particles is raised because phenolic species are known as
during the ignition of the wood (Elsasser et al., 2013), which is concur- antioxidants and scavengers of reactive oxygen species. Thus phenolic
rently the part of the combustion with the highest variation and might species could hypothetically mitigate the overall toxicity of harmful
give an explanation of this finding. It was observed that logs of birch ig- substances (Kjällstrand and Petersson, 2001).
nited faster than the other two logwoods and showed the lowest For atmospheric studies and a comparison between the investigated
amount of residues in the chamber before the introduction of the con- stoves with stoves of older design, the relative amount of phenolic spe-
secutive batch, which supports the explanation. However, the percent- cies to other particle constituents is more important than the absolute
age of levoglucosan to the total particle mass in the presented quantity. Therefore, the total phenolic content of the particles was nor-
experiments was generally very low. Although total PM2.5 mass was malised to OC. Beech, birch and pellet combustion featured similar, but
not determined, it can be estimated by assuming that TC represents ap- significantly higher contents of phenolic species per OC (3.7 mg/gOC,
proximately 50% of PM2.5. Hence, levoglucosan accounts for b 0.3% of 3.4 mg/gOC and 3.9 mg/gOC) than spruce (1.7 mg/gOC). In comparison
PM2.5 for all particle emissions from the masonry heater, which is with combustion appliances without air-staging (Table 3), absolute
very low compared to the literature data (Orasche et al., 2012; emissions of particle-bound phenolic species and phenolic species per
Schmidl et al., 2008; Vicente et al., 2015a) and might be a feature of OC from both masonry heater and pellet boiler were one order of mag-
the applied combustion technology and associated higher temperatures nitude lower. Although different combustion scenarios are considered,
in the firebox. In this context, only the pellet boiler led to comparable re- the comparison can be regarded as conservative because only the hot ig-
sults to previous studies with 2.7% levoglucosan in PM (Orasche et al., nition of three consecutive batches, known for lower emissions of phe-
2012; Vicente et al., 2015b). However, please note that Vicente et al. nolic species than cold starts (Orasche et al., 2012), was taken into
(2015a), Vicente et al. (2015b) and Schmidl et al. (2008) sampled account, whereas the combustion in this study included six consecutive
PM10 instead of PM2.5 as in this study which might affect the results batches and a char-burning phase, but a cold start as well. Moreover, the
as single analytes can vary over size fractions of combustion particles analysed 23 phenolic species were the most abundant ones in this
(Kleeman et al., 2008). Although wood combustion PM2.5 from the study. Therefore, the emissions of these two stoves with air-staging sec-
two appliances of this study even appear solely in the fraction of PM1, ondary air supply represent a different type of emission pattern, which
it is not known for stoves of previous studies, so a direct comparison becomes even more relevant since combustion technologies in small-
was omitted. Nevertheless, the obtained ratios of levoglucosan to OC, scale appliances are currently developing and old stoves are constantly
quantified from PM2.5, of 9% for beech logwood (Calvo et al., 2015) replaced by more efficient modern stoves.
and 19% and 10% for beech and spruce logwood (Orasche et al., 2012)
from conventional stoves were substantially higher than for the modern 3.3.4. Resin acids and phytosterols
masonry heater (beech: 0.58%; birch: 0.88%; spruce: 1.18%), which sup- The detected resin acids comprise derivatives and thermal degrada-
ports our findings. tion products of abietic acid and its isomers pimaric and isopimaric acid.
The EFs of the further quantified anhydrous sugars, mannosan and These resin compounds are known to be mainly released during the
galactosan, were one order of magnitude lower than levoglucosan in start phase of the combustion of coniferous wood (Orasche et al.,
all experiments. Mannosan is released from the decomposition of hemi- 2012). Of all resin compounds, dehydroabietic acid was the most abun-
cellulose, thus the ratio of levoglucosan to mannosan (LGS/MNS) has dant compound with a mean EF of 42.12 μg MJ−1 for spruce wood, but
been proposed to differentiate between hardwood and softwood com- partially oxidised derivatives and products from elimination reactions
bustion. Hard- and softwoods contain different amounts of cellulose of abietic acid were present as well. Small amounts of dehydroabietic
and hemicellulose, which is also reflected in the emission. LGS/MNS of acid were also found on particles from birch wood combustion (8.56
15.0 ± 7.1 for beech, 19.3 ± 13.5 for birch and 8.5 ± 2.4 for spruce re- μg MJ−1), which raise suspicion that these EFs were caused by carry-
veal the same trend as previously reported, but with higher standard over effects. However, this EF results from two of four birch wood com-
deviations (Schmidl et al., 2008). In addition to that, this concept does bustion experiments which were not carried out directly after spruce
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 641

wood. Furthermore, a porous-tube dilutor, which was cleaned every German Research Foundation as described in Orasche et al. (2012). Fi-
day, was used in the sampling instead of a dilution tunnel which was nally, the TEQs of single PAHs were summed up to the total PAH-TEQ
proposed to minimise carry-over effects (Orasche et al., 2012). There- for a single combustion experiment.
fore, we regard the detected quantities as truly abundant in the birch PAH-TEQ EFs of logwood combustion range from 0.57 μg MJ−1 to 1.4
particulate emissions. μg MJ−1 without significant differences between the investigated wood
In addition to spruce, one characteristic substance was found for types. Similar to the PAH-EFs, PAH-TEQ EFs of the pellet boiler were al-
each of the investigated wood fuels. β-Sitosterol was only abundant in most two orders of magnitude lower than the logwoods of this study
the particulate emissions of the pellet boiler (10.4 μg MJ−1) in two of with values from 0.019 μg MJ−1 to 0.034 μg MJ−1. These findings are
three experiments. According to the manufacturer, the main material in line with other small-scale wood stoves with advanced combustion
of the pellets was spruce and pine wood, but concentrations of β- technology and more than one order of magnitude lower than conven-
sitosterol occurred below the limit of quantification in every single com- tional stoves (Fig. 2) and agree with the general trend of decreasing
bustion experiment of spruce logwood. Substantial quantities of this TEQ-EFs for automatically-fired wood combustion appliances. However,
triterpenoid were measured on particles from the combustion of differ- the concept of PAH-TEQ only considers the abundance of PAHs and does
ent pine species (Oros and Simoneit, 2001a), thus the emission of β- not include other harmful compounds or even easing synergetic effects,
sitosterol is likely associated with content of pine wood of the pellets. such as the influence of antioxidants.
Lupa-2,22(29)-dien-28-ol refers to an thermal alteration product of
betulin, which is one of the major components in essential oil of birch 3.3.6. PAH-derivatives
wood, especially in its bark (Laszczyk et al., 2006). Its EF accounts for In this section, aromatic hydrocarbons with heteroatoms within or
29.0 μg MJ− 1 for birch wood, but some smaller quantities could be at the periphery of the core structure are discussed. The most abundant
also detected for spruce wood. hydroxyl- and oxy-PAHs (OH-PAHs and OPAHs) were denoted by 1,8-
For beech, substantial amounts of stigmasta-3,5-dien-7-one, an al- naphthalic anhydride, 9H-fluoren-9-one, 1-hydroxynaphthalene and
teration product of stigmastan, was found with an EF of 173.90 9,10-anthracenedione. All OH\\ and OPAHs were significantly lower
μg MJ−1. Minor amounts (5.41 μg MJ−1) were also detected for birch, in the pellet boiler particle emissions (1.90 μg MJ−1 in total) or appeared
which belongs to deciduous trees as well as beech, that this stigmastan below the limit of quantification, whereas differences between the log-
and derivatives were previously proposed as marker for this type of wood were less pronounced. Only for spruce significantly elevated EFs
woods (Gonçalves et al., 2011). were observed for 7H-benzo[c]fluoren-7-one, 11H-benzo[b]fluoren-
It must be added that the analysed resin acids and triterpenoids only 11-one and 7H-benzo[de]anthracen-7-one compared to birch. Never-
represent a minor part of the total alteration products of abietic acid, theless, the highest EFs for total OH\\ and OPAH were obtained for
betulin and stigmastan (Oros and Simoneit, 2001a; Oros and Simoneit, beech (63.4 μg MJ−1) with almost equal EFs for birch (40.9 μg MJ−1)
2001b). On that account, marker substances for specific wood species and spruce (40.4 μg MJ−1). Dibenzothiophene was the only sulphur-
or classes cannot be derived. containing polycyclic aromatic compound (PASH) found on the parti-
cles with substantial increased EFs for birch wood.
3.3.5. Parent and alkylated PAH Compared to conventional stoves and pellet boilers, the total emis-
sion of OH\\ and OPAH was decreased by more than one order of
3.3.5.1. Emission profiles and factors. In all combustion experiments,
phenanthrene, pyrene and fluoranthene were the most abundant
PAHs and accounted from 48.1% (spruce) up to 73.0% (pellets) of the
total analysed PAHs. However, the gas-particle partitioning of especially
these PAHs is sensitive to even small changes in sampling temperature,
which could have affected the presented EFs for particle-bound PAHs.
Although significant differences could be found between the logwoods,
the combustion of spruce logs slightly shifted the emission pattern to-
wards PAHs of higher molecular weight, which might be caused by py-
rolysis of resin components. Additionally, a higher degree of alkylation
can be observed for spruce wood which is relevant for the development
of diagnostic ratios for source apportionment (Tobiszewski and
Namieśnik, 2012). A more detailed discussion on aromatic profiles can
be found in Section 3.4.
The EF of total PAHs from the pellet boiler (0.89 μg MJ−1) was more
than one order of magnitude lower than from the masonry heater
fuelled with any logwood (beech: 35.31 μg MJ− 1; birch: 32.31
μg MJ−1; spruce: 22.16 μg MJ−1), which emphasises the efficiency of
this combustion appliance. Generally, the EFs of particle-bound PAH
were one to two orders of magnitude lower compared to conventional
wood stoves and pellet boilers (Orasche et al., 2012), but similar to
other modern masonry heaters (Tschamber et al., 2016), which also
agrees with the trend of volatile aromatic emissions investigated in pre-
vious studies (Czech et al., 2017; Czech et al., 2016).
Fig. 2. EFs of PAH-TEQ values for different wood combustion appliances. The dashed line
separates results of this study (left) from literature data (aTschamber et al., 2016;
3.3.5.2. PAH toxicity equivalents (PAH-TEQ). PAHs adsorbed on particu- b
Orasche et al., 2012) (right). WABI and XP54-IN denote wood stoves with modern
late matter are well-known for their carcinogenic and mutagenic activ- combustion technology, while pellet and LS (logwood stove) are conventional wood
ity (Lewtas, 2007). On that account, the approach of PAH toxicity combustion appliances at cold start (CS) or nominal load (NL). Error bars represent
equivalent (PAH-TEQ) was developed to assess the health risk potential minimum and maximum PAH-TEQ of repeated experiments. The results demonstrate
that modern technology for batchwise wood combustion not only reduces EFs by one
at workplace exposures by multiplication of single PAH concentrations order of magnitude, but also the toxicological potential of the emissions related to the
or EFs with its corresponding toxicity equivalent factor (TEF). TEFs are same heating values of the wood fuels. PAH-TEQ EFs for the pellet boiler are even
related to benzo[a]pyrene which accounts for 1 and taken from the reduced by an additional order of magnitude.
642 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

magnitude. Additionally, their distribution was shifted towards com- poor in peak number and intensity and therefore were not considered
pounds of lower molecular weight for the same type of fuel wood further.
(Orasche et al., 2012). Therefore, the differences in OH\\and OPAH pat- In the preceding sections, the discussion of the two experiments
terns might be explained by the level of combustion technology. with slow ignition, spruce* and birch*, were omitted because of differ-
ences in aromatic patterns. With all combustion experiments, a hierar-
3.4. Untargeted analysis of particle-bound aromatic compounds by TOCA- chical cluster analysis was performed based on unweighted pair group
REMPI-TOFMS method with arithmetic means (Sokal and Rohlf, 1962) with coefficient
of congruence as distance metric (Abdi, 2010). Three clusters were ob-
An approach for the rapid characterisation of aromatic compounds tained and assigned to spruce, pellets and the hardwoods birch and
in thermal fractions of particulate matter offers the hyphenation of beech (Fig. 4). Despite the correct assignment of spruce*, its distance
resonance-enhanced multi-photon ionisation time-of-flight mass spec- was remarkably larger than between the other spruce combustion ex-
trometry (REMPI-TOFMS) to thermal/optical carbon analysis (TOCA). periments. Moreover, birch* appeared isolated from the other birch
The first two OC-fractions defined by the ImproveA protocol for carbon combustion experiments closer related to the pellets. Both spruce*
analysis were combined because it is assumed that organic particle con- and birch* emitted higher amounts of OGC during the first of the six
stituents predominantly evaporate without decomposition. In OC3 and batches, which prompts an inappropriate or slow ignition. In a previous
OC4, pyrolysis occurs which shifts the mass spectrum towards lower m/ study, the molecular composition of the VOCs during the first batch of
z (Diab et al., 2015). spruce* revealed higher quantities of typical primary decomposition
In Fig. 4, mean REMPI mass spectra for each wood fuel are illus- products from lignin or carbohydrates, e.g. coniferyl alcohol and
trated with labels of the most important homologue alkylation series levoglucosenone, as it can be observed when combustion efficiency de-
of phenanthrene (m/z 178), pyrene (m/z 202), benzo[a]anthraces/ creases (Czech et al., 2016). Due to the fact that these combustion
chrysene (m/z 228) and dibenzofuran (m/z 168) with m/z incre- events of a relatively short period remarkably affect the emissions of
ments of + 14. Although response factor (photoionisation cross sec- the entire 4 h experiment, spruce* and birch are separately discussed
tions) can vary substantially (Boesl, 2000), peak intensities reflected in the following section.
the relations between the wood fuels described by the quantitative
results from IDTD-GCMS. 3.5. Effect of slow ignition
Highest peak intensities were observed for phenanthrenes and
pyrenes with its respective maximum at the parent PAH and a steady 3.5.1. Spruce wood
decay towards higher degrees of alkylation, which was also found for The REMPI spectrum of spruce* with slow ignition showed a higher
particles from other wood stoves (Bente et al., 2008; Diab et al., 2015) number of peaks, a broader m/z range and a different pattern than the
and volatile PAH emissions as well (Elsasser et al., 2013). However, a spruce wood experiments with proper ignition (Fig. 5, top). Parent
slight increase from methyl- to C2-phenanthrene (sum of all isomers PAHs were less dominating the homologue series so that the general de-
of dimethyl- and ethyl-phenanthrene) can be observed in beech-, gree of alkylation is higher than in the regular combustion experiments.
birch- and pellet-derived particles, which is possibly an issue of mass Moreover, larger PAHs, such as the homologue series of benzopyrenes
interference with another isobaric compound of m/z 206, e.g. and benzoperylenes, appeared with higher abundances. Many peaks
phenanthrene-9-carboxaldehyd. Furthermore, the thermal degradation were located between the signals of PAH homologue series and likely
of abietic acid from spruce wood leads to the formation of retene (m/z belong to monomers of the lignin decomposition, which are increasing-
234), but to lower alkylated phenanthrenes as well (e.g. pimanthrene ly released at lower combustion temperature, i.e. during smouldering
(Simoneit, 2002)), so the findings from other PAH homologue series (Kjällstrand and Petersson, 2001). Additionally, abietic acid (m/z 302)
cannot be transferred to phenanthrenes. In contrast to distinct undergoes aromatisation of its saturated rings through thermal alter-
abundancies of phenolic species in PM2.5 of conventional wood stoves ation and become accessible for REMPI, for example as dihydroretene
(Diab et al., 2015), the absence of such m/z supports the results from or simonellite (Oros and Simoneit, 2001a). Peaks of m/z N 302 may be
IDTD-GC/MS measurements and the general low emissions of the mod- assigned to lignans or larger PAHs as discussed in a previous section.
ern masonry heater. Generally, the observed higher intensities for spruce* compared to the
In the m/z range of the spectra, which can be regarded as fingerprint regular experiments were confirmed by the higher EFs for several or-
region for the wood types, peaks up to m/z 400 were still detected. Al- ganic compounds from IDTD-GCMS analysis.
though reliable assignments to molecular structures were not possible, With the EFs of Table 2, a Dean-Dixon outlier test (at α = 0.05)
the number of potential formula can be reduced by the constraints that was performed to identify significantly increased or decreased sub-
the detected compound must be an aromatic constituent of wood stances for slow ignition. Dean-Dixon test was chosen because it
smoke due to the ionisation selectivity. Hence, it is hypothesised that does not assume normal distribution of the data and is generally rec-
the peaks in the m/z range from 300 to 400 (small panels on Fig. 3) be- ommended for small sample sizes (Dean and Dixon, 1951). In
long to lignans, i.e. dimers of monomers derived from lignin, or alter- addition to previously mentioned enhancement of OGC, a variety
ation products of steroids and triterpenoids, which may become of PAHs, OH\\ and OPAHs, among others benzo[a]anthracene,
aromatised under thermal stress and thus become accessible for benzo[a]pyrene and 7H-benzo[de]anthracen-7-one, were identified
REMPI. The latter proposed compound class seems to be more likely be- as outliers with increased EFs. Therefore, it was not surprising that
cause substantially more peaks were detected for birch wood combus- the PAH-TEQ EF were changed by a factor of six (8.82 μg MJ− 1) com-
tion. In contrast to the other wood fuels, birch contains high amounts pared to the regular spruce combustion (1.44 μg MJ− 1 ). Further-
of triterpenoids, such as betulin or betulinic acid, in its bark (Laszczyk more, primary decomposition products from lignin, such as vanillic
et al., 2006), which leads to a variety of possible thermal alteration prod- acid, coniferyl aldehyde and guaiacylacetone, were more prominent
ucts (Oros and Simoneit, 2001b) and may justify the higher number of in spruce*, in agreement with the REMPI mass spectra, as well as the
peaks in the upper m/z range. anhydrous sugars levoglucosan and mannosan. Regarding inorganic
In the pyrolysis fraction OC3 (Fig. S7), a shift towards lower m/z, in- emissions, spruce* was the only spruce combustion experiment in
cluding the appearance of volatile aromatic compounds such as naph- which chromium was found above the limit of quantification. How-
thalene or phenol, could be observed as expected because of the ever, on the basis of the presented data this phenomenon could not
decomposition of larger molecular structures. Apart from retene as an be explained, but since chromium does not belongs to volatile metals
indicator for the combustion of coniferous wood, no specific peaks or ar- in wood combustion, contamination is assumed. Further inorganic
omatic patterns were found. The obtained mass spectra of OC4 were particle constituents were not affected by the slow ignition.
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 643

Fig. 3. Mean REMPI mass spectra of summed intensities of OC1 + OC2 for the four investigated wood types. Symbols denote the most abundant series of parent PAHs and alkylated
homologues, namely dibenzofuran (green triangles), phenanthrene (red squares), pyrene (cyan stars) and chrysene/benzo[a]anthracene (blue circles). Panels in the upper right corner
belong to fingerprint regions of the wood types, which contain alteration products of phytosterols and lignans and enable wood type identification despite similar alkylation patterns.
Figures with logarithmic y-scale can be found in the Supplementary data (Fig. S8).

Out of nine quantified components or alteration products of the


spruce wood resin, four (isopimaric acid, abietic acid, dehydroabietic
acid and dehydroabietic acid methyl ester) were found as outliers
with higher abundance than in the regular spruce wood combustion
and one (7-oxo-abietic acid) with lower abundance. Isopimaric acid
and abietic acid were even not detected in each of the regular spruce
combustion experiments, but with EFs of 73.27 μg MJ−1 and 435.58
μg MJ−1 in spruce*, suggesting an enhanced distillation-like release dur-
ing the start phase of the first batch (Orasche et al., 2012). In contrast,
the absence of 7-oxo-abietic acid in spruce* implies its formation during
higher temperatures sufficient to partially oxidise abietic acid. Due to
the fact that pimaric acid, a primary resin constituent as well, were
only found in the regular spruce combustion experiments, a possible
heat-induced isomerisation between pimaric, isopimaric and abietic
acid might occur. All together, these results emphasise the necessary de-
tection of many possible resin acids and their alteration products if
quantitative conclusion are drawn in source apportionment studies
since the abundance of a single markers strongly depends on the com-
bustion condition.

3.5.2. Birch wood


Fig. 4. Dendrogram from cluster analysis of REMPI mass spectra from OC1 + OC2 using the
unweighted pair group method with arithmetic means and coefficient of congruence as
Although birch* was classified as slow ignition experiment as well, it
distance metric. Experiments spruce* and birch* were regarded as outliers and separately exhibited a different emission pattern than spruce* (Fig. 5, bottom). In
discussed in the Section 3.5. this case, the slow ignition caused a distinct shift towards higher m/z
644 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

Fig. 5. Mean REMPI mass spectra of summed intensities of OC1 + OC2 for combustion experiment with slow ignition. Symbols denote the homologue series of dibenzofuran (green up-
pointing triangles), phenanthrene (red squares), pyrene (cyan stars), chrysene/benzo[a]anthracene (blue circles), benzpyrene (magenta diamonds) and benzo[ghi]perylene (black down-
pointing triangles). Panels in the upper right corner belong to fingerprint regions of the wood fuels, which contain lignans and alteration products of phytosterols. Although both
combustion experiments reveal a shift towards larger PAH emissions, the spectrum of spruce* contains numerous peaks, possibly from thermal resin alteration, while for birch*
emissions parent PAHs are more pronounced in the alkylation series. Figures with logarithmic y-scale can be found in the Supplementary data (Fig. S9).

and an elevated contribution of parent PAHs to their homologue series. inorganic constituents, which differ qualitatively and quantitatively
Moreover, pyrene (m/z 202) replaced phenanthrene (m/z 178) as the from the investigated masonry heater emissions. For both modern ma-
most abundant PAH. However, apart from PAH homologue series the sonry heater and pellet boiler, wood-specific combustion products from
number of peaks and peak intensities were low, indicating that primary the decomposition of lignin and carbohydrates were detected, but sub-
decomposition products from lignin were not or less elevated in this stantially lower compared to conventional wood stoves and pellet
combustion scenario. Therefore, birch* was located closer to the pellet boilers. Although the major single component of organics in the PM2.5
boiler particles in the cluster analysis (Fig. 4). was levoglucosan, its contribution to OC was found at the lower limit
Also in this case the findings from the REMPI mass spectra agree described in the literature, which also holds for phenolic species. Con-
with the results obtained from IDTD-GCMS analysis and Dean-Dixon versely, the higher relative amount of general products of incomplete
test. 25 PAHs, among others the carcinogenic benzo[a]pyrene, combustion, such as parent PAHs and OPAHs, complicates the assign-
dibenz[ah]anthracene and anthanthrene, and seven alkylated PAHs ment of this type of emissions to wood combustion in atmospheric
were found to be statistical outliers compared to the regular birch com- source apportionment studies. This holds especially for CMB modelling
bustion experiments, which accompany the 14-fold increased PAH-TEQ and simpler approaches, such as the estimation of wood smoke mass
(birch*: 11.3 μg MJ−1; mean regular birch: 0.82 μg MJ−1). Additionally, from ambient levoglucosan concentrations (Schmidl et al., 2008), but
12 OPAHs with two to five rings were classified as outliers as well with also for PMF factor identification as tracers for wood combustion are
fold changes between 5 and 34, which is more pronounced than for substantially reduced (Bari et al., 2009). This might become a more im-
spruce*. The majority of the 12 OPAHs and 25 PAHs were found to cor- portant issue in the future because of the rather long period of use for
relate well and significantly with the formation of reactive oxygen spe- wood stoves and boiler. Especially the pellet boiler lacks wood combus-
cies (Sklorz et al., 2007) and therefore OPAHs are of toxicological tion markers, but β-sitosterol was found to be only component which
interest although they are not considered in the TEQ concept. was more abundant in the pellet boiler emissions than in the modern
With the Dean-Dixon test primary decomposition products were not masonry heater emissions. However, β-sitosterol cannot be regarded
identified as outliers except syringol, so birch* deals with a different in- as marker for pellet emissions because of its high dependency on the
fluence of the slow ignition than spruce*. However, it could not be con- wood type. Nevertheless, our results provide comprehensive emission
cluded whether the different emission patterns belong to features of the profiles of modern wood combustion appliances which can be already
wood species or combustion conditions because of the number of used in chemical mass balances for areas of northern Europe.
experiments. With the hyphenation of thermal/optical carbon analysis to
resonance-enhanced multi-photon ionisation time-of-flight mass spec-
4. Conclusion trometry (TOCA-REMPI-TOFMS), differences in aromatic fingerprints
and PAH alkylation series of the different PM2.5 samples were
In this comprehensive study of wood combustion PM2.5 from two discussed. Moreover, it was shown that slow ignition of the first batch
modern small-scale combustion appliances (masonry heater and pellet in batchwise logwood combustion increased EFs of single species in dif-
boiler) equipped with air staging, several emission factors for organic ferent manners and also shifted the aromatic pattern to larger struc-
and inorganic particle constituents were presented and compared to tures. Slow ignition could be linked to elevated PAH-TEQ values as
other state-of-the-art and conventional wood stoves. Generally, mason- well as different quantities of phytosterols and the distribution of their
ry heater emissions appeared in the same order of magnitude as for a thermal alteration products, which are used as wood-specific markers.
woods stoves equipped with primary, secondary and post-combustion Finally, emission source apportionment should be connected to
air (Tschamber et al., 2016), but more than one order of magnitude more sophisticated toxicological approaches instead of simple concepts
lower than the conventional wood stove (Orasche et al., 2012). Howev- of PAH-TEQ. Due to the presence of substantial amounts of antioxidants
er, the pellet boiler even reduces carbonaceous emissions by an addi- in the wood smoke which may counteract harmful effects on human
tional order of magnitude. Pellet boiler PM2.5 is mainly comprised of health by single particle constituents, further studies under realistic
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 645

Table 2
Mean emission factors (EFs) and concentrations of gaseous analytes, particulate carbons, inorganic elements, PAHs, PAH-derivatives, phenolic species, resin constituents and sterols as well
as toxicity equivalents (TEQ) and modified combustion efficiency (MCE) from regular combustion experiments (modern masonry heater: beech, birch and spruce; pellet boiler: softwood
pellets) and combustion experiments with slow ignition (modern masonry heater: spruce* and birch*; see Section 3.5). Minimum and maximum EFs are listed in Table S2. EFs can be
related to the amount of burned fuel by using respective net heating values from Table 1. Entries of b.l.q. refer to EFs below the limit of quantification.

Unit Beech Birch Spruce Pellet Spruce* Birch*

Number of experiments n 4 4 3 3 1 1

Gaseous analytes
O2 % 16.8 16.0 15.9 12.0 16.1 15.5
CO2 % 3.91 4.55 4.38 8.55 4.48 4.56
CO mg MJ−1 1320 1430 1640 42.4 1290a 1230
OGC mg MJ−1 24.0 15.4 20.1 1.25 45.0a 18.6
NOx mg MJ−1 110 104 83.0 89.0 83.1 99.9
MCE [] 0.987 0.988 0.983 0.999 0.986a 0.984a

Particulate carbon
OC mg MJ−1 20 6.1 6.1 0.15 11 5.7
EC mg MJ−1 27 47 13 0.27 17 34
TC mg MJ−1 47.0 53.1 19.1 0.42 28.0 39.7

Inorganic particle constituents


Al μg MJ−1 45.1 82.5 39.1 22.3 38.5 30.4
B μg MJ−1 3.64 5.45 4.35 5.31 b.l.q. 3.92
Ba μg MJ−1 1.63 1.67 1.94 1.16 1.40 1.17
Bi μg MJ−1 b.l.q. b.l.q. b.l.q. 15.74 b.l.q. b.l.q.
Ca μg MJ−1 83.7 89.1 101.7 24.2 51.3 109
Cd μg MJ−1 0.97 2.49 0.22 b.l.q. b.l.q. 1.59
Cr μg MJ−1 2.42 3.78 b.l.q. 1.68 8.38a 2.17
Cu μg MJ−1 15.5 6.36 4.18 7.65 6.52 11.7
Fe μg MJ−1 20.5 38.2 29.0 12.1 21.8 18.3
K μg MJ−1 1630 2050 1600 3010 1120 1440
Li μg MJ−1 b.l.q. b.l.q. b.l.q. 4.95 b.l.q. b.l.q.
Mn μg MJ−1 4.75 7.51 7.41 9.31 4.80 5.99
Mo μg MJ−1 b.l.q. b.l.q. b.l.q. 0.84 b.l.q. 1.40a
Na μg MJ−1 51.2 52.6 53.7 924 b.l.q. 49.1
Pb μg MJ−1 9.80 8.89 b.l.q. 1.21 b.l.q. 6.44
S μg MJ−1 393 370 337 754 301 266
Ti μg MJ−1 0.90 1.62 1.16 0.47 1.03 1.87
W μg MJ−1 b.l.q. 13.0 3.76 2.80 7.44 9.49
Zn μg MJ−1 53.5 683 302 139 269 513

Particle-bound PAH
Phenanthrene μg MJ−1 9.34 11.4 2.62 0.25 1.30 16.6
Anthracene μg MJ−1 1.27 1.28 0.42 0.02 b.l.q. b.l.q.
Fluoranthene μg MJ−1 7.65 6.38 4.47 0.22 3.40 20.6a
Acephenanthrylene μg MJ−1 0.63 0.71 0.33 0.01 1.16 4.08a
Pyrene μg MJ−1 6.74 5.28 3.56 0.18 4.62 20.1a
Benzo[a]fluorene μg MJ−1 0.43 0.14 0.46 0.01 1.52 3.15a
Benzo[b]fluorene μg MJ−1 0.43 0.20 0.55 0.01 1.33 1.69a
Benzo[c]phenanthrene μg MJ−1 0.37 0.41 0.34 0.01 1.23a 2.33a
Benzo[ghi]fluoranthene μg MJ−1 1.78 1.69 1.38 0.03 4.31 10.1a
Benz[a]anthracene μg MJ−1 0.75 0.55 0.83 0.01 4.73a 5.99a
Cyclopenta[cd]pyrene μg MJ−1 b.l.q. 0.10 b.l.q. b.l.q. 6.75a 2.71a
Chrysene μg MJ−1 1.38 1.14 1.46 0.03 5.00a 10.2a
2,2′-Binaphthalene μg MJ−1 b.l.q. b.l.q. 0.09 b.l.q. b.l.q. 0.48a
sum Benzo[b,j,k]fluoranthene μg MJ−1 2.02 1.31 2.24 0.04 11.6a 14.2a
Benz[e]pyrene μg MJ−1 0.78 0.48 0.88 0.02 3.86a 5.68a
Benz[a]pyrene μg MJ−1 0.88 0.55 0.97 0.02 5.81a 7.77a
Perylene μg MJ−1 0.03 0.02 0.10 b.l.q. 1.01a 1.07a
Indeno[1,2,3-cd]fluoranthene μg MJ−1 b.l.q. 0.24 b.l.q. b.l.q. 0.12a 0.06
Benzo[b]triphenylene μg MJ−1 b.l.q. b.l.q. 0.01 b.l.q. 0.10a 0.07a
Indeno[7,1,2,3-cdef]chrysene μg MJ−1 b.l.q. b.l.q. 0.04 b.l.q. 0.57a 0.33a
1-Phenylpyrene μg MJ−1 b.l.q. b.l.q. 0.01 b.l.q. b.l.q. 0.10a
Dibenz[ah]anthracene μg MJ−1 b.l.q. b.l.q. 0.04 b.l.q. 0.23 0.46a
Indeno[1,2,3-cd]pyrene μg MJ−1 0.33 0.18 0.48 b.l.q. 1.75 3.6a
Benzo[b]chrysene μg MJ−1 b.l.q. b.l.q. 0.01 b.l.q. b.l.q. 0.08a
Pentaphene μg MJ−1 b.l.q. 0.04 0.01 b.l.q. b.l.q. 0.06
Picene μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. 0.31a 0.04a
Benzo[ghi]perylene μg MJ−1 0.50 0.18 0.69 0.02 3.23 5.51a
Anthanthrene μg MJ−1 b.l.q. b.l.q. 0.02 b.l.q. 0.50a 0.40a
Coronene μg MJ−1 b.l.q. b.l.q. 0.15 0.01 1.3a 1.9a
Sum PAH μg MJ−1 35.3 32.4 22.2 0.9 65.8 139.3a
PAH-TEQ μg MJ−1 1.26 0.818 1.44 0.028 8.82a 11.3a

Particle-bound OH-PAH, OPAH & PASH


1-Hydroxynaphthalene μg MJ−1 7.5 1.5 1.7 1.0 b.l.q. 2.5
2-Hydroxynaphthalene μg MJ−1 0.38 0.61 0.44 b.l.q. b.l.q. 2.8
1,8-Dihydroxynaphthalene μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. 1.4a b.l.q.

(continued on next page)


646 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

Table 2 (continued)

Unit Beech Birch Spruce Pellet Spruce* Birch*

Number of experiments n 4 4 3 3 1 1

1-Naphthalenecarboxaldehyde μg MJ−1 b.l.q. 0.10 b.l.q. b.l.q. b.l.q. 0.26


2-Naphthalenecarboxaldehyde μg MJ−1 0.08 0.35 b.l.q. b.l.q. b.l.q. 0.55
1(2H)-Acenaphthylenone μg MJ−1 0.07 0.07 0.05 b.l.q. 0.02 0.28
9H-Fluoren-9-one μg MJ−1 6.9 10 5.1 0.30 0.88 17
1H-Phenalen-1-one μg MJ−1 3.3 2.2 3.6 0.07 20a 20a
Xanthone μg MJ−1 3.5 2.0 2.8 0.08 3.9 6.0
9,10-Anthracenedione μg MJ−1 3.9 3.1 5.2 0.10 6.6 17a
Benzo[b]naphtho[1,2-d]furan μg MJ−1 1.8 0.22 0.44 0.19 b.l.q.a 0.43a
Cyclopenta(def)phenanthrenone μg MJ−1 3.5 2.5 3.6 0.08 3.4 17
1,8-Naphthalic anhydride μg MJ−1 27 15 8.0 b.l.q. 50a 49.7
Benzo[b]naphtho[2,1-d]furan μg MJ−1 1.2 0.46 0.92 0.02 b.l.q. 4.6a
2,3-5,6-Dibenzoxalene μg MJ−1 0.35 0.19 0.37 0.01 b.l.q. 1.1
Benzo[b]naphtho[2,3-d]furan μg MJ−1 0.37 0.11 0.37 0.01 b.l.q. 1.7
Benzo[kl]xanthene μg MJ−1 0.33 0.16 0.38 b.l.q. b.l.q. 1.6a
11H-Benzo[a]fluoren-11-one μg MJ−1 0.20 0.07 0.43 0.01 1.4 1.4a
7H-Benzo[c]fluorene-7-one μg MJ−1 0.09 0.03 0.14 b.l.q. b.l.q. 0.54a
11H-Benzo[b]fluoren-11-one μg MJ−1 0.45 0.11 1.0 0.01 2.8 3.8a
7H-Benzo[de]anthracen-7-one μg MJ−1 0.83 0.24 1.4 0.01 5.0a 7.2a
Naphtho[2,1,8,7-klmn]xanthene μg MJ−1 0.46 0.54 0.55 0.01 b.l.q. 1.8
Benz[a]anthracene-7,12-dione μg MJ−1 b.l.q. 0.15 0.6 b.l.q. 0.31 1.1
5,12-Naphthacenedione μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q. 0.21a
Phenanthro[3,4-c]furan-1,3-dione μg MJ−1 0.09 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
6H-Benzo[cd]pyren-6-on μg MJ−1 1.8 0.73 3.3 0.02 3.0 15a
Dibenzothiophene μg MJ−1 0.01 0.06 0.01 b.l.q. b.l.q. 0.13

Particle-bound alkylated PAH


9-Methylphenanthrene μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q. 0.19a
3,6-Dimethylphenanthrene μg MJ−1 0.02 0.22 0.51 b.l.q. b.l.q. 0.37
Retene μg MJ−1 b.l.q. 0.20 6.5 0.04 11a b.l.q.
Sum 2-/8-methylfluoranthene μg MJ−1 0.45 0.19 0.48 0.01 0.64 2.0a
Sum 1-/3-/7-methylfluoranthene μg MJ−1 0.62 0.28 0.64 0.02 1.0 2.9a
4-Methylpyrene μg MJ−1 0.22 0.07 0.28 b.l.q. 0.62 2.0a
2-Methylpyrene μg MJ−1 0.59 0.31 0.69 0.02 1.2 2.5a
1-Methylpyrene μg MJ−1 0.46 0.23 0.52 0.02 1.9a 2.4a
4,5-Dimethylpyrene μg MJ−1 0.06 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
2,3,6,7-Tetramethylanthracene μg MJ−1 b.l.q. b.l.q. 0.35 b.l.q. b.l.q. b.l.q.
1-Methyl-benz[a]anthracene μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q. 0.53a

Particle-bound anhydrous sugars


Galactosan μg MJ−1 2.3 2.2 2.9 b.l.q. 15 5.9
Mannosan μg MJ−1 7.9 3.8 9.2 0.47 150a 6.7
Levoglucosan μg MJ−1 110 53 72 10 1100a 56

Particle-bound phenolic species


4-Hydroxyphenylethanol μg MJ−1 1.3 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
4-Hydroxybenzoic acid μg MJ−1 4.0 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Vanillin μg MJ−1 10.1 14.0 5.94 0.37 13.7 11.6
Isoeugenol μg MJ−1 2.45 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Acetovanillone μg MJ−1 2.1 0.5 0.8 b.l.q. b.l.q. 0.61
Methylvanillate μg MJ−1 1.1 b.l.q. 0.27 0.01 b.l.q.a b.l.q.
Vanillic acid μg MJ−1 0.37 1.5 b.l.q. b.l.q. 16a 0.8
3-Guaiacylpropanol μg MJ−1 0.60 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Coniferaldehyde μg MJ−1 1.30 b.l.q. 0.80 b.l.q. 470a b.l.q.
Guaiacylacetone μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. 0.63a b.l.q.
4-Guaiacylbutanoic acid μg MJ−1 0.57 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Syringol μg MJ−1 1.74 0.04 b.l.q. b.l.q. b.l.q. 0.46a
4-Methylsyringol μg MJ−1 1.11 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
4-Ethylsyringol μg MJ−1 1.19 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Allylsyringol μg MJ−1 0.08 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Syringaldehyde μg MJ−1 36 3.7 0.44 0.09 b.l.q. 1.9
Syringylpropene μg MJ−1 0.71 0.13 0.02 0.01 b.l.q. 0.06
Acetosyringone μg MJ−1 4.4 0.29 0.03 b.l.q. b.l.q. 0.21
Syringic acid, methyl ester μg MJ−1 0.91 0.04 b.l.q. b.l.q. b.l.q. 0.03
Syringic acid μg MJ−1 0.91 0.17 0.10 0.11 0.50a 0.12
Homosyringic acid μg MJ−1 0.12 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Sinapylaldehyde μg MJ−1 0.62 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
3,5-Dimethoxy-4,4′-dihydroxystilbene μg MJ−1 0.95 b.l.q. b.l.q. b.l.q. b.l.q. b.l.q.
Sum phenolic species μg MJ−1 72.1 20.4 8.42 0.59 504a 15.8

Particle-bound resin constituents


Pimaric acid μg MJ−1 b.l.q. b.l.q. 0.06 b.l.q. b.l.q. b.l.q.
Isopimaric acid μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. 73a b.l.q.
Abietic acid μg MJ−1 b.l.q. b.l.q. b.l.q. b.l.q. 440a b.l.q.
6-Dehydrodehydroabietic acid μg MJ−1 b.l.q. b.l.q. 1.1 b.l.q. b.l.q. b.l.q.
Dehydroabietic acid μg MJ−1 b.l.q. 8.6 42 b.l.q. 710a b.l.q.
7-Oxodehydroabietic acid μg MJ−1 b.l.q. b.l.q. 0.08 b.l.q. b.l.q.a b.l.q.
H. Czech et al. / Science of the Total Environment 612 (2018) 636–648 647

Table 2 (continued)

Unit Beech Birch Spruce Pellet Spruce* Birch*

Number of experiments n 4 4 3 3 1 1

6-Dehydrodehydroabietic acid, methyl ester μg MJ−1 b.l.q. b.l.q. 0.09 b.l.q. b.l.q. b.l.q.
Dehydroabietic acid, methyl ester μg MJ−1 b.l.q. 0.16 0.82 0.01 7.33a b.l.q.
7-Oxodehydroabietic acid, methyl ester μg MJ−1 b.l.q. b.l.q. 0.09 b.l.q. b.l.q. b.l.q.

Particle-bound phytosterols
β-Sitosterol μg MJ−1 b.l.q. b.l.q. b.l.q. 10 b.l.q. b.l.q.
Stigmasta-3,5-dien-7-one μg MJ−1 170 5.4 b.l.q. b.l.q. b.l.q. b.l.q.
Lupa-2,22(29)-dien-28-ol μg MJ−1 b.l.q. 29 4.6 b.l.q. b.l.q. b.l.q.
a
Significant outlier from Dean-Dixon test at 0.05 significance level in comparison to regular combustion experiments.

Table 3
Emission factor of total phenolic species and their normalisation to OC from modern masonry heater (beech, birch and spruce logs) and pellet boiler (softwood pellets) compared to con-
ventional logwood and pellet combustion from Orasche et al. (2012).

Beech Birch Spruce Pellet

μg mg/gOC μg mg/gOC μg mg/gOC μg mg/gOC


MJ−1 MJ−1 MJ−1 MJ−1

This study 72 3.7 20 3.4 8.4 1.4 0.59 3.4


Orasche et al., 2012a 2220 150 – – 690 62 18 26
a
“PB spruce”, “LS beech” and “LS spruce”.

exposure conditions, e.g. with lung cells in an air-liquid interface Chow, J.C., Watson, J.G., Chen, L.W.A., Chang, M.C.O., Robinson, N.F., Trimble, D., et al.,
2007. The IMPROVE_A temperature protocol for thermal/optical carbon analysis:
(Kanashova et al., 2017; Paur et al., 2011), may support a full assessment maintaining consistency with a long-term database. J. Air Waste Manage. Assoc. 57,
of improvements in stove construction as well as the influence of com- 1014–1023.
Croft, D.P., Cameron, S.J., Morrell, C.N., Lowenstein, C.J., Ling, F., Zareba, W., et al., 2017. As-
bustion condition and wood type. sociations between ambient wood smoke and other particulate pollutants and bio-
markers of systemic inflammation, coagulation and thrombosis in cardiac patients.
Acknowledgement Environ. Res. 154, 352–361.
Czech, H., Sippula, O., Kortelainen, M., Tissari, J., Radischat, C., Passig, J., et al., 2016. On-line
analysis of organic emissions from residential wood combustion with single-photon
The measurements were carried out at the University of Eastern ionisation time-of-flight mass spectrometry (SPI-TOFMS). Fuel 177, 334–342.
Czech, H., Pieber, S.M., Tiitta, P., Sippula, O., Kortelainen, M., Lamberg, H., et al., 2017.
Finland (UEF), Department of Environmental and Biological Science, in
Time-resolved analysis of primary volatile emissions and secondary aerosol forma-
cooperation with the Helmholtz Virtual Institute of Complex Molecular tion potential from a small-scale pellet boiler. Atmos. Environ. 158, 236–245.
Systems in Environmental Health (HICE), funded by The Helmholtz Im- Dean, R.B., Dixon, W.J., 1951. Simplified statistics for small numbers of observations. Anal.
Chem. 23, 636–638.
pulse and Network Fund of the Helmholtz Association (Germany), the Diab, J., Streibel, T., Cavalli, F., Lee, S.C., Saathoff, H., Mamakos, A., et al., 2015. Hyphenation
DACH-project WooShi (grant ZI 764/5-1), the Academy of Finland of a EC/OC thermal-optical carbon analyzer to photo-ionization time-of-flight mass
(grants 296645 and 304459) and the University of Eastern Finland for spectrometry: an off-line aerosol mass spectrometric approach for characterization
of primary and secondary particulate matter. Atmos. Meas. Tech. 8, 3337–3353.
the project “sustainable bioenergy, climate change and health”. Elsasser, M., Busch, C., Orasche, J., Schön, C., Hartmann, H., Schnelle-Kreis, J., et al., 2013.
Dynamic changes of the aerosol composition and concentration during different
burning phases of wood combustion. Energy Fuel 27, 4959–4968.
Appendix A. Supplementary data Evtyugina, M., Alves, C., Calvo, A., Nunes, T., Tarelho, L., Duarte, M., et al., 2014. VOC emis-
sions from residential combustion of Southern and mid-European woods. Atmos. En-
The supplementary data involves the ICP-OES method, minimum viron. 83, 90–98.
Fountoukis, C., Butler, T., Lawrence, M.G., Denier van der Gon, H.A.C., Visschedijk, A.J.H.,
and maximum EF for experimental repetitions of the same conditions, Charalampidis, P., et al., 2014. Impacts of controlling biomass burning emissions on
volcano diagrams of ANOVA post-hoc tests, REMPI mass spectra of wintertime carbonaceous aerosol in Europe. Atmos. Environ. 87, 175–182.
OC3 and REMPI spectra of OC1 + OC2 with logarithmic y-scale. Supple- Fuller, G.W., Sciare, J., Lutz, M., Moukhtar, S., Wagener, S., 2013. New directions: time to
tackle urban wood burning? Atmos. Environ. 68, 295–296.
mentary data associated with this article can be found in the online ver- Gonçalves, C., Alves, C., Fernandes, A.P., Monteiro, C., Tarelho, L., Evtyugina, M., et al., 2011.
sion, at http://dx.doi.org/10.1016/j.scitotenv.2017.08.263. Organic compounds in PM2.5 emitted from fireplace and woodstove combustion of
typical Portuguese wood species. Atmos. Environ. 45, 4533–4545.
Grabowsky, J., Streibel, T., Sklorz, M., Chow, J.C., Watson, J.G., Mamakos, A., et al., 2011. Hy-
References phenation of a carbon analyzer to photo-ionization mass spectrometry to unravel the
organic composition of particulate matter on a molecular level. Anal. Bioanal. Chem.
Abdi, H., 2010. Congruence: congruence coefficient, Rv-coefficient and Mantel coefficient. 401, 3153–3164.
In: Salkind, N. (Ed.), Encyclopedia of Research Design. Sage Publications, Thousand Hellén, H., Hakola, H., Haaparanta, S., Pietarila, H., Kauhaniemi, M., 2008. Influence of res-
Oaks, CA. idential wood combustion on local air quality. Sci. Total Environ. 393, 283–290.
Andreae, M.O., Ramanathan, V., 2013. Climate's dark forcings. Science 340, 280–281. Jalava, P.I., Happo, M.S., Kelz, J., Brunner, T., Hakulinen, P., Mäki-Paakkanen, J., et al., 2012.
Bari, M.A., Baumbach, G., Kuch, B., Scheffknecht, G., 2009. Wood smoke as a source of In vitro toxicological characterization of particulate emissions from residential bio-
particle-phase organic compounds in residential areas. Atmos. Environ. 43, mass heating systems based on old and new technologies. Atmos. Environ. 50, 24–35.
4722–4732. Kanashova T, Sippula O, Oeder S, Streibel T, Passig J, Czech H, et al. Emissions from a mod-
Bente, M., Sklorz, M., Streibel, T., Zimmermann, R., 2008. Online laser desorption- ern log wood masonry heater and wood pellet boiler: composition and biological im-
multiphoton postionization mass spectrometry of individual aerosol particles: molec- pact on air-liquid interface exposed human lung cells. (manuscript in preparation)
ular source indicators for particles emitted from different traffic-related and wood 2017.
combustion sources. Anal. Chem. 80, 8991–9004. Kari, E., Hao, L., Yli-Pirilä, P., Leskinen, A., Kortelainen, M., Grigonyte, J., et al., 2017. Effect
Boesl, U., 2000. Laser mass spectrometry for environmental and industrial chemical trace of pellet boiler exhaust on secondary organic aerosol formation from α-pinene. Envi-
analysis. J. Mass Spectrom. 35, 289–304. ron. Sci. Technol. 51, 1423–1432.
Bruns, E.A., El Haddad, I., Slowik, J.G., Kilic, D., Klein, F., Baltensperger, U., et al., 2016. Iden- Kasurinen, S., Jalava, P.I., Happo, M.S., Sippula, O., Uski, O., Koponen, H., et al., 2017. Partic-
tification of significant precursor gases of secondary organic aerosols from residential ulate emissions from the combustion of birch, beech, and spruce logs cause different
wood combustion. Sci Rep 6, 27881. cytotoxic responses in A549 cells. Environ. Toxicol. 32, 1487–1499.
Calvo, A.I., Martins, V., Nunes, T., Duarte, M., Hillamo, R., Teinilä, K., et al., 2015. Residential Khodaei, H., Guzzomi, F., Patiño, D., Rashidian, B., Yeoh, G.H., 2017. Air staging strategies
wood combustion in two domestic devices: relationship of different parameters in biomass combustion-gaseous and particulate emission reduction potentials. Fuel
throughout the combustion cycle. Atmos. Environ. 116, 72–82. Process. Technol. 157, 29–41.
648 H. Czech et al. / Science of the Total Environment 612 (2018) 636–648

Kjällstrand, J., Petersson, G., 2001. Phenolic antioxidants in wood smoke. Sci. Total Envi- Reda, A.A., Czech, H., Schnelle-Kreis, J., Sippula, O., Orasche, J., Weggler, B., et al., 2015.
ron. 277, 69–75. Analysis of gas phase carbonyl compounds in emissions from modern wood combus-
Kleeman, M.J., Robert, M.A., Riddle, S.G., Fine, P.M., Hays, M.D., Schauer, J.J., et al., 2008. tion appliances: influence of wood type and combustion appliance. Energy Fuel 29,
Size distribution of trace organic species emitted from biomass combustion and 3897–3907.
meat charbroiling. Atmos. Environ. 42, 3059–3075. Schmidl, C., Marr, I.L., Caseiro, A., Kotianová, P., Berner, A., Bauer, H., et al., 2008. Chemical
Künzi, L., Mertes, P., Schneider, S., Jeannet, N., Menzi, C., Dommen, J., et al., 2013. Re- characterisation of fine particle emissions from wood stove combustion of common
sponses of lung cells to realistic exposure of primary and aged carbonaceous aerosols. woods growing in mid-European Alpine regions. Atmos. Environ. 42, 126–141.
Atmos. Environ. 68, 143–150. Schmidl, C., Luisser, M., Padouvas, E., Lasselsberger, L., Rzaca, M., Ramirez-Santa Cruz, C., et
Lamberg, H., Nuutinen, K., Tissari, J., Ruusunen, J., Yli-Pirilä, P., Sippula, O., et al., 2011a. al., 2011. Particulate and gaseous emissions from manually and automatically fired
Physicochemical characterization of fine particles from small-scale wood combus- small scale combustion systems. Atmos. Environ. 45, 7443–7454.
tion. Atmos. Environ. 45, 7635–7643. Sehlstedt, M., Dove, R., Boman, C., Pagels, J., Swietlicki, E., Löndahl, J., et al., 2010. Antiox-
Lamberg, H., Sippula, O., Tissari, J., Jokiniemi, J., 2011b. Effects of air staging and load on idant airway responses following experimental exposure to wood smoke in man.
fine-particle and gaseous emissions from a small-scale pellet boiler. Energy Fuel 25, Part. Fibre Toxicol. 7.
4952–4960. Shafizadeh, F., 1968. In: Tipson, MLWaRS (Ed.), Pyrolysis and Combustion of Cellulosic
Laszczyk, M., Jäger, S., Simon-Haarhaus, B., Scheffler, A., Schempp, C.M., 2006. Physical, Materials. Advances in Carbohydrate Chemistry 23, pp. 419–474.
chemical and pharmacological characterization of a new oleogel-forming triterpene Simoneit, B.R.T., 2002. Biomass burning - a review of organic tracers for smoke from in-
extract from the outer bark of birch (Betulae cortex). Planta Med. 72, 1389–1395. complete combustion. Appl. Geochem. 17, 129–162.
Lewtas, J., 2007. Air pollution combustion emissions: characterization of causative agents Sippula, O., Hytönen, K., Tissari, J., Raunemaa, T., Jokiniemi, J., 2007. Effect of wood fuel on
and mechanisms associated with cancer, reproductive, and cardiovascular effects. the emissions from a top-feed pellet stove. Energy Fuel 21, 1151–1160.
Mutat. Res. Rev. Mutat. Res. 636, 95–133. Sippula, O., Lind, T., Jokiniemi, J., 2008. Effects of chlorine and sulphur on particle forma-
Lipsky, E.M., Robinson, A.L., 2006. Effects of dilution on fine particle mass and partitioning tion in wood combustion performed in a laboratory scale reactor. Fuel 87,
of semivolatile organics in diesel exhaust and wood smoke. Environ. Sci. Technol. 40 2425–2436.
(1), 155–162. Sippula, O., Lamberg, H., Leskinen, J., Tissari, J., Jokiniemi, J., 2017. Emissions and ash be-
Martinsson, J., Eriksson, A.C., Nielsen, I.E., Malmborg, V.B., Ahlberg, E., Andersen, C., et al., havior in a 500 kW pellet boiler operated with various blends of woody biomass
2015. Impacts of combustion conditions and photochemical processing on the light and peat. Fuel 202, 144–153.
absorption of biomass combustion aerosol. Environ. Sci. Technol. 49, 14663–14671. Sklorz, M., Briedé, J.J., Schnelle-Kreis, J., Liu, Y., Cyrys, J., De Kok, T.M., et al., 2007. Concen-
Miljevic, B., Heringa, M.F., Keller, A., Meyer, N.K., Good, J., Lauber, A., et al., 2010. Oxidative tration of oxygenated polycyclic aromatic hydrocarbons and oxygen free radical for-
potential of logwood and pellet burning particles assessed by a novel profluorescent mation from urban particulate matter. J. Toxicol. Environ. Health A Curr. Issues 70,
nitroxide probe. Environ. Sci. Technol. 44, 6601–6607. 1866–1869.
Naeher, L.P., Brauer, M., Lipsett, M., Zelikoff, J.T., Simpson, C.D., Koenig, J.Q., et al., 2007. Sokal, R.R., Rohlf, F.J., 1962. The comparison of dendrograms by objective methods. Taxon
Woodsmoke health effects: a review. Inhal. Toxicol. 19, 67–106. 11, 33–40.
Nordin, E.Z., Uski, O., Nyström, R., Jalava, P., Eriksson, A.C., Genberg, J., et al., 2015. Influ- Sommersacher, P., Brunner, T., Obernberger, I., 2012. Fuel indexes: a novel method for the
ence of ozone initiated processing on the toxicity of aerosol particles from small evaluation of relevant combustion properties of new biomass fuels. Energy Fuel 26,
scale wood combustion. Atmos. Environ. 102, 282–289. 380–390.
Nussbaumer, T., 2003. Combustion and co-combustion of biomass: fundamentals, tech- Tiitta, P., Leskinen, A., Hao, L., Yli-Pirilä, P., Kortelainen, M., Grigonyte, J., et al., 2016. Trans-
nologies, and primary measures for emission reduction. Energy Fuel 17, 1510–1521. formation of logwood combustion emissions in a smog chamber: formation of sec-
Nuutinen, K., Jokiniemi, J., Sippula, O., Lamberg, H., Sutinen, J., Horttanainen, P., et al., ondary organic aerosol and changes in the primary organic aerosol upon daytime
2014. Effect of air staging on fine particle, dust and gaseous emissions from masonry and nighttime aging. Atmos. Chem. Phys. 16, 13251–13269.
heaters. Biomass Bioenergy 67, 167–178. Tobiszewski, M., Namieśnik, J., 2012. PAH diagnostic ratios for the identification of pollu-
Orasche, J., Schnelle-Kreis, J., Abbaszade, G., Zimmermann, R., 2011. Technical note: in-situ tion emission sources. Environ. Pollut. 162, 110–119.
derivatization thermal desorption GC-TOFMS for direct analysis of particle-bound Tschamber, V., Trouvé, G., Leyssens, G., Le-Dreff-Lorimier, C., Jaffrezo, J.L., Genevray, P., et
non-polar and polar organic species. Atmos. Chem. Phys. 11, 8977–8993. al., 2016. Domestic wood heating appliances with environmental high performance:
Orasche, J., Seidel, T., Hartmann, H., Schnelle-Kreis, J., Chow, J.C., Ruppert, H., et al., 2012. chemical composition of emission and correlations between emission factors and op-
Comparison of emissions from wood combustion. Part 1: emission factors and char- erating conditions. Energy Fuel 30, 7241–7255.
acteristics from different small-scale residential heating appliances considering par- Uski, O., Jalava, P.I., Happo, M.S., Torvela, T., Leskinen, J., Mäki-Paakkanen, J., et al., 2015.
ticulate matter and polycyclic aromatic hydrocarbon (PAH)-related toxicological Effect of fuel zinc content on toxicological responses of particulate matter from pellet
potential of particle-bound organic species. Energy Fuel 26, 6695–6704. combustion in vitro. Sci. Total Environ. 511, 331–340.
Oros, D.R., Simoneit, B.R.T., 2001a. Identification and emission factors of molecular tracers Vicente, E.D., Duarte, M.A., Calvo, A.I., Nunes, T.F., Tarelho, L.A.C., Custódio, D., et al., 2015a.
in organic aerosols from biomass burning part 1. Temperate climate conifers. Appl. Influence of operating conditions on chemical composition of particulate matter
Geochem. 16, 1513–1544. emissions from residential combustion. Atmos. Res. 166, 92–100.
Oros, D.R., Simoneit, B.R.T., 2001b. Identification and emission factors of molecular tracers Vicente, E.D., Duarte, M.A., Tarelho, L.A.C., Nunes, T.F., Amato, F., Querol, X., et al., 2015b.
in organic aerosols from biomass burning part 2. Deciduous trees. Appl. Geochem. 16, Particulate and gaseous emissions from the combustion of different biofuels in a pel-
1545–1565. let stove. Atmos. Environ. 120, 15–27.
Pastorello, C., Caserini, S., Galante, S., Dilara, P., Galletti, F., 2011. Importance of activity Watson, J.G., Chow, J.C., Houck, J.E., 2001. PM2.5 chemical source profiles for vehicle ex-
data for improving the residential wood combustion emission inventory at regional haust, vegetative burning, geological material, and coal burning in Northwestern Col-
level. Atmos. Environ. 45, 2869–2876. orado during 1995. Chemosphere 43, 1141–1151.
Paur, H.R., Cassee, F.R., Teeguarden, J., Fissan, H., Diabate, S., Aufderheide, M., et al., 2011. Weggler, B.A., Ly-Verdu, S., Jennerwein, M., Sippula, O., Reda, A.A., Orasche, J., et al., 2016.
In-vitro cell exposure studies for the assessment of nanoparticle toxicity in the lung-a Untargeted identification of wood type-specific markers in particulate matter from
dialog between aerosol science and biology. J. Aerosol Sci. 42, 668–692. wood combustion. Environ. Sci. Technol. 50, 10073–10081.
Pio, C., Cerqueira, M., Harrison, R.M., Nunes, T., Mirante, F., Alves, C., et al., 2011. OC/EC Wiinikka, H., Grönberg, C., Boman, C., 2013. Emissions of heavy metals during fixed-bed
ratio observations in Europe: re-thinking the approach for apportionment between combustion of six biomass fuels. Energy Fuel 27, 1073–1080.
primary and secondary organic carbon. Atmos. Environ. 45, 6121–6132.

You might also like