Owned by: China Association for Science and Technology
Sponsored by: China Coal Society
Published by: Springer Nature
About issue

The International Journal of Coal Science & Technology is a peer-reviewed open access journal. It focuses on key topics of coal scientific research and mining development, serving as a forum for scientists to present research findings and discuss challenging issues.


Coverage includes original research articles, new developments, case studies and critical reviews in all aspects of scientific and engineering research on coal, coal utilizations and coal mining. Among the broad topics receiving attention are coal geology, geochemistry, geophysics, mineralogy, and petrology; coal mining theory, technology and engineering; coal processing, utilization and conversion; coal mining environment and reclamation and related aspects.


The International Journal of Coal Science & Technology is published with China Coal Society, who also cover the publication costs so authors do not need to pay an article-processing charge.


The journal operates a single-blind peer-review system, where the reviewers are aware of the names and affiliations of the authors, but the reviewer reports provided to authors are anonymous.


  • A forum for new research findings, case studies and discussion of important challenges in coal science and mining development

  • Offers an international perspective on coal geology, coal mining, technology and engineering, coal processing, utilization and conversion, coal mining environment and reclamation and more

  • Published with the China Coal Society

Show More
Editors-in-Chief
Suping Peng, Shimin Liu
Managing Editor
Wanjie Wang
Associate Editors
Bo Hyun Kim, Dongjie Xue, Pedram Roghanchi, Wu Xiao, Zhiqiang Wu
Publishing model
Open Access. Learn about publishing OA with us
Home > Volumes and issues > Volume 9, issue 1

Chemistry and particle size distribution of respirable coal dust in underground mines in Central Eastern Europe

Research Article

Open Access

Published: 20 February 2022

0 Accesses

International Journal of Coal Science & Technology Volume 9, article number 3, (2022)

Abstract

Despite international efforts to limit worker exposure to coal dust, it continues to impact the health of thousands of miners across Europe. Airborne coal dust has been studied to improve risk models and its control to protect workers. Particle size distribution analyses shows that using spraying systems to suppress airborne dusts can reduce particulate matter concentrations and that coals with higher ash yields produce finer dust. There are marked chemical differences between parent coals and relatively coarse deposited dusts (up to 500 µm, DD500). Enrichments in Ca, K, Ba, Se, Pb, Cr, Mo, Ni and especially As, Sn, Cu, Zn and Sb in the finest respirable dust fractions could originate from: (i) mechanical machinery wear; (ii) variations in coal mineralogy; (iii) coal fly ash used in shotcrete, and carbonates used to reduce the risk of explosions. Unusual enrichments in Ca in mine dusts are attributed to the use of such concrete, and elevated K to raised levels of phyllosilicate mineral matter. Sulphur concentrations are higher in the parent coal than in the DD500, probably due to relatively lower levels of organic matter. Mass concentrations of all elements observed in this study remained below occupational exposure limits.

1.Introduction

Around 66% and 57% of current European coal production and demand respectively is restricted to central-eastern countries such as Poland, Germany, Czech Republic and Ukraine (BP 2020). In this region, existing coal reserves comprise a total of 53,383 Mt of anthracite and bituminous reserves and 56,945 Mt of sub-bituminous and lignite reserves, which together account for 82% of European Union and around 11% of worldwide coal reserves (BP 2020). Given the continued importance of this industry in eastern Europe, and although coal mining activity in Europe has been steadily decreasing during the last few years (Jonek-Kowalska 2018), the environmental impact of coal mining on workers’ health remains a matter of concern (Więckol-Ryk et al. 2018; Song et al. 2020). Of particular relevance is the amount and chemistry of the coal dust produced in underground coal mines. Coal dust is a chemically heterogeneous mixture of carbonaceous matter, with a variable mineral content (e.g. quartz, sulphates and sulphides), and has distinctive major (Fe, Si, S) and trace element (As, Pb, Cr, Mn, U) signatures which may be linked to classic coal worker diseases such as pneumoconiosis (CWP) or progressive massive fibrosis (PMF) (Dalal et al. 1995; Liu et al. 2005; Cohen et al. 2008; Li et al. 2013; Caballero-Gallardo and Olivero-Verbel 2016; Fang et al. 2020; Pedroso-Fidelis et al. 2020; Jin et al. 2021). In recent years, the use of improved technologies has generally increased mine output but has also, in some cases, produced higher concentrations of dust (Perret et al. 2017; Johann-Essex et al. 2017; Leonard et al. 2020; Fan and Liu 2021). This may have been a factor in the resurgence of coal mining-related respiratory diseases in some mines (e.g. Fan et al., 2018; Wu et al., 2019). For this reason, dust monitoring control during the handling of different mining technologies requires extensive investigation, including information on the physicochemical characteristics of the deeply inhalable (particle size below 10 µm) and respirable fractions (particle size below 4 µm and 2.5 µm), both of which could be a key factor in the prevention of coal dust problems in mining, such as diverse authors are contributing (Zhang et al. 2020; Szkudlarek and Janas 2021; Jiang and Luo 2021; Liu et al. 2021; Reed et al. 2021).

In this study, as part of the European ROCD project (Reducing risks from Occupational exposure to Coal Dust), we focussed on the chemical characterisation of coarse (< 500 µm) and respirable (< 4 and < 2.5 µm) coal dusts and their parent coals from underground mines in Central-Eastern Europe. This included determining the nature and origins of chemical fractionation of major and trace elements between different size fractions of dust. In addition, samples of mine airborne particulate matter below 10 µm (PM10), emitted during different coal mining activities, were also obtained and analysed to evaluate possible chemical differences and the efficiency of ventilation and air suppression systems.

2.Geological settings

A total of 16 underground coal mines have been sampled for this study (Table S1 in Supplementary Material), all of them located in two coal mining areas of Central-Eastern Europe: Upper Silesian Coal Basin (USCB, 15 mines) and Velenje Coal Basin (VCB, 1 mine).

2.1 Upper silesian coal basin

The USCB, which is one of the major coal basins in Europe, covers an area of around 7500 km2 (5800 km2 in Poland and 1600 km2 in the Czech Republic, Fig. 1a); only 30% of its exploitable coal resources have been worked (Kotarba 2001; Kusiak et al. 2006; Stępniewska et al. 2014; Vaněk et al. 2017). It is a basin of polygenetic origin: the lower part (Namurian A) consists of an Upper Carboniferous coal-bearing lithostratigraphic sequence deposited in a paralic environment, whereas the upper part (Namurian B to Wesphalian D) is of purely continental origin (Kotas and Porzycki 1984; Kotas et al. 1994; Kotarba 2001; Kusiak et al. 2006).

Fig. 1
figure 1

a Upper: Geological map of the Upper Silesian Coal Basin in Southern Poland. A–A’ indicates the location of the geological cross section (lower figure).

The USCB hosts a wide range of coal types, from low-rank (sub-bituminous), high- to low-volatile rank (bituminous) to high-rank (anthracite) (Kotarba 2001; Kusiak et al. 2006; Kędzior 2009). Only the bituminous and sub-bituminous coals are mined and used in power stations (EPA 1995; Kedzior 2015). On average, USCB coals contain 0.9%–2% sulphur, have an ash yield of 11%–16% and calorific values ranging from 29 to 32 MJ/kg (EPA 1995).

2.2 Velenje coal basin

The VCB is located in NE Slovenia, at the SE end of the Alps, 70 km NE of Ljubljana (Fig. 1b). The coal has an average calorific value of 8–13 MJ/kg, an ash yield of 5%–40% db, a moisture content of 15%–45% and sulphur content of 0.5%–2% db (Markič and Sachsenhofer 2010). The lignite-bearing Pliocene to Early Quaternary sediments accumulated in a 12 km long and 4 km wide WNW-ESE trending, fault bounded, depression. The mainly clastic basin fill is up to 1000 m thick, and contains a lenticular 10 km long, 3 km wide and up to 160 m thick lignite seam of Pliocene age (Brezigar 1987; Markic and Sachsenhofer 1997; Kędzior 2019).

3.Methodology

3.1 Mine sampling

A total of 91 deposited dust (DD) samples were collected manually inside the mines using two different methods. The most common method involved using a brush and a hermetic-sterilised plastic bag. Additionally, some coal DD samples were collected, using a plastic tray left in a specific workplace within the underground mine for 24–48 h, before being stored in a sterilised container. These two different coal dust collection methods offered the opportunity to compare relatively “aged” dust (which accumulated over an unknown time period, collected using a brush) with fresh DD on trays. In addition, given the likelihood that dust-generating activities carried out in the underground mines is related to the coal seam being worked, 16 rock samples of ‘parent’ coals were collected (Table S2 in Supplementary Material). See sampling details in Tables S3, S4 and S5 in Supplementary Material. Parent coals refer to coal samples extracted from the coal seam exploited during the DD sampling collection. They were used to examine any possible change in chemical composition compared to the DD samples.

3.2 Sample analysis

3.2.1 Sample pre-treatment

Figure 2 provides an overview of the analytical procedures undertaken: (i) DD samples were sieved to < 500 µm (DD500) for the determination of particle size distribution (PSD), proximate, ultimate and geochemical composition; (ii) a small (riffle-split) portion (100 g) of these DD samples (as received) was size separated using a special PM4 or PM2.5 separator/collection device to produce a respirable DD fraction (RDD, dust finer than 4 µm or 2.5 µm, RDD4 or RDD2.5, respectively); (iii) parent coal rock samples were homogenised by crushing (using a hammer), mixing, and milling (using a Mortar Agatha Grinder RM 200 device in an automatic mode), prior to proximate and geochemical analysis.

Fig. 2
figure 2

Schematic of the sampling procedures used for parent coal, deposited dust and separation of the respirable deposited dust, and subsequent analyses

In detail, to obtain the RDD4 or RDD2.5 samples, a separator was used following the protocols previously detailed by Amato et al. (2009a, b) and Moreno et al. (2019). The sieved coal DD500 samples were placed within the separator (which comprises a methacrylate chamber which has a volume of 4697 cm2 by 17 cm of diameter) and subjected to mild resuspension as a result of slow rotation (1.5 rpm). The resuspended particles were sampled by drawing air, at a rate of 5 or 25 L/min for RDD2.5 and RDD4 respectively, through a size-selective inlet head which deposits PM4 or PM2.5 on to a 0.60 μm pore size, 47 mm diameter polycarbonate filter. This type of filter is strong enough, does not contaminate the sample, and allows for a relatively easy extraction from its surface. After that, the RDD samples were ready for chemical analysis.

3.2.2 Particle size distribution

The PSD of DD and RDD samples was determined using a Malvern Mastersizer device. For DD samples, a Scirocco 2000 unit was coupled to this device in order to disperse and separate the particulate in an air stream ready for laser diffraction size determination. For RDD samples (with very low sample volumes) PSD was obtained using a wet (water-polyphosphate RDD suspension) HydroG 2000 coupled unit to verify their RDD4 size separation (Sperazza et al. 2004).

3.2.3 Proximate, ultimate and chemical characterization

Coal sample proximate and ultimate analyses were done following ISO and ASTM procedures (ISO-589, 1981, ISO-1171, 1976, ISO-562, 1974, ASTM D-3286, D-3302 M, D3174-12), with moisture (M) and ash yields (HTA) obtained at 105 and 750 °C.

Prior to geochemical analysis, coal, DD500 and RDD samples were HF-HNO3-HClO4 acid digested following the method of Querol et al. (1997, 1992) to retain potentially volatile elements, such as As and Se. The resulting sample solutions were analysed for major and trace elements by Inductively-Coupled Plasma Atomic-Emission Spectrometry for major elements (ICP-AES, Iris Advantage Radial ER/S device from Thermo Jarrell-Ash) and Inductively-Coupled Plasma Mass Spectrometry for trace elements (ICP-MS, X-SERIES II Thermo Fisher Scientific, Bremen, Germany). International reference materials SARM19 and NIST SRM 1633b, and blanks were treated in the same way. After thirteen analyses of the above reference materials, analytical errors and relative standard deviations (RSDVs) were included in Table S6 in Supplementary Material.

3.2.4 PM10 measurements

The use of most electrical PM monitoring and collection devices in underground coal mines is not possible due to the risk of coal dust explosion. For this reason, only ATEX-certified instruments are permitted. In this study, a TECORA-CIP10 thoracic PM sampler (CIP10), designed for operation in explosive environments, was used to determine PM10 ambient concentrations during specific coal mining activities. The sampler, which operates at a flow rate of 10 L/min, has a PM10-selective inlet that retains the > 10 µm PM on an impaction surface and collects the PM10 onto a polyurethane foam.

The lack of amount sample retained in the filter foam makes sometimes impossible for their digestions, for this reason, some filter collection keeping always same conditions were occasionally combined. The mass of airborne PM10 per volume of air is calculated from the weight of the two fractions divided by the volume sampled. The CIP10 samplers were mainly operated for a period of 80–135 min, or, in a few cases, for 60–160 min.

The PM10 fractions retained in the polyurethane foam were subject to acid digestion using a mixture of HF, HNO3 and HClO4, following the method previously discussed in the Sect. 3.2.3. The whole foam was digested. However, in this case 10 mL of HNO3 was used in the first step, instead of 2.5 mL, and then all of the HNO3 was evaporated off to eliminate the organic excess part of the foam (polyurethane) and avoid combustion in the second acid digestion step. Finally, all sample digested was recovered in the final of the second step ready for ICP-MS and ICP-AES analysis. Blank foams were digested using the same protocol and their contained concentrations of different elements then subtracted from the measured concentrations in the samples.

4.Results and discussion

4.1 Particle size distribution

A total of 91 coal DD samples from both coal basins and 2 Dust samples from VCB warehouse (Malta and Calplex, dust samples used in some places of the underground mine for security, preventing ignitions or collapses) were analysed for their PSD, moisture content (%M, air dried, ad) and ash yield (%HTA, dry bases, db) (Tables S7 and S8 in Supplementary Material). Mean contents of the < 500 (DD500), < 10 (DD10), < 4 (DD4), and < 2.5 µm (DD2.5) in DD from both coal basins reached 98.0%, 17.0%, 6.0% and 3.3% volume (%v), respectively, for VCB samples, and 98%, 35%, 16% and 10%v, respectively, for USCB. Thus, the proportions of DD10, DD4 and DD2.5 in the DD samples from USCB were more than double those from the VCB. In contrast, mean moisture contents were higher in the VCB samples (12%ad VCB vs. 1.0%ad USCB), and ash yield means were similar (36% for VCB and 38% for USCB).

The markedly higher moisture and lower DD10, DD4 and DD2.5 proportions in the VCB DD500 are probably due to wetting of airborne and deposited coal dust in the underground mines and not to differences in coal rank. If that is the case, then it is likely that it has reduced markedly the proportion of finer particle sizes by agglomeration of particles. Thus, in the VCB underground mine, PSD analysis of very low moisture dust samples yield DD2.5 fractions reaching up to 23% DD500. The efficiency of wetting DD in reducing coal mine worker exposure to dust, as well as in preventing coal dust explosions, is very well documented (Woskoboenko 1988; Küçük et al. 2003; Kuai et al. 2012; Yuan et al. 2014; Ajrash et al. 2017; Azam et al. 2019; Hu et al. 2021; Zhang et al. 2021). Küçük et al. (2003) and Yuan et al. (2014) found that a decrease in the dust fraction below < 125 µm can be achieved by increasing the moisture content of deposited coal dust. These studies emphasise the importance of keeping a minimum moisture content in DD to prevent explosions. There is a balance to be found, however, as maintaining relatively low ambient humidity levels in underground mines provides a more comfortable working environment (Sunkpal et al. 2017) which improves work-place productivity.

When moisture contents of DD500 and then DD10, DD4 and DD2.5 were determined independently for each of the two basins, some significant correlations were found for samples from the VCB mine (Eqs. (1)–(9)); no such correlation was found in the USCB samples. This is probably because the moisture contents are very low (0.2%ad to 5.4%ad) in USCB, compared to VCB (5.5%ad to 27%ad) samples. On the other hand, systematic correlations between size fraction and HTA yields were only significant in the USCB DD samples, where a wider range of HTA yields of DD500 were obtained (8.2%ad to 79%ad in USCB, compared with 19%ad to 47%ad in VCB).

Another plausible explanation to account for the negative correlation between moisture content and fine DD500 fractions, different to the abatement of the fine dust fraction caused by the wetting of dust, is that the moisture content in DD500 from VCB increases with the coal content of the dust. An elevated coal matrix content in dust can yield to coarser dust, as well as higher moisture content (because this a lignite and accordingly moisture increases with the increase of the organic content), without the need of increasing moisture content by wetting DD. However, we do not believe this is the case because of the lack of correlation between ash yields and the fine fractions of DD500 in VCB. Thus, we point to the different degrees of wetting of the DD500 samples to account for the different concentrations of the fine fractions. On the other hand, this wetting may be hiding possible correlations between ash yields and fine fraction sizes, as found for the low moisture dust from USCB.

In the case of the DD collected at the VCB mine, these included fresh DD deposited on a plastic tray (24–48 h collection time, 9 samples) and aged DD brushed from surfaces (9 samples). There was therefore a total of 18 of the 21 samples analysed (3 samples were considered outliers due to problems in their collection or store). Results for the different methods of collection are shown below to evaluate possible differences in the PSD of DD samples collected with the two methods.

Velenje coal basin samples

Fresh DD samples (n = 9/10)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = - 0.{285}\% {{M}},{\text{ad}} + {7}.{328} \;R^{{2}} = 0.{746}$$
(1)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = - 0.{4}0{7 }\% {{M}},{\text{ad}} + {11}.{887} \;{R}^{{2}} = 0.{677}$$
(2)
$${\text{DD}}_{10} \left( {\% {\text{v}}} \right) = - 1.087 \, \% {{M}},{\text{ad}} + 32.201 \;R^{2} = 0.667$$
(3)

Aged DD samples (n = 9/11)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = - 0.{229 }\% {{M}},{\text{ad}} + {6}.{17}0 \;{R}^{{2}} = 0.{779}$$
(4)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = - 0.{329 }\% {{M}},{\text{ad}} + {1}0.{176}\; {R}^{{2}} = 0.{7}0{7}$$
(5)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = - 0.{9}0{5 }\% {{M}},{\text{ad}} + {28}.{276}\;{R}^{{2}} = 0.{639}$$
(6)

All DD samples (n = 18/21)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = - 0.{25}0\% {{M}},{\text{ad}} + {6}.{68}0 \;{R}^{2} = 0.{754}$$
(7)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = - 0.{359 }\% {\text{M}},{\text{ad}} + {1}0.{948}\;{R}^{{2}} = 0.{688}$$
(8)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = - 0.{975 }\% {{M}},{\text{ad}} + {3}0.0{3}0\;{R}^{{2}} = 0.{648}$$
(9)

Very similar results were obtained for the two types of DD samples, showing that the DD sampling protocols collect dust with similar PSDs. Also, when the two warehouse samples are added, the correlation between moisture and particle size increases, presumably because warehouse samples (Malta and Calplex) have much lower moisture contents and much higher HTA yields (0.4%ad and 0.6%ad moisture and 98 and 62%db HTA), and therefore contain more fine dust.

Upper silesian coal basin samples

Marcel DD samples (n = 8/10)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = 0.{495 }\% {\text{HTA}},{\text{db}} - {9}.0{75} \;{R}^{{2}} = 0.{958}$$
(10)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = 0.{664 }\% {\text{HTA}},\;{\text{db}} - {1}0.{416} \;{R}^{{2}} = 0.{977}$$
(11)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = 0.{949 }\% {\text{HTA}},{\text{db}} - {1}0.{5}0{2}\;{R}^{{2}} = 0.{982}$$
(12)

Bielszowice DD samples (n = 16/16)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = 0.{422 }\% {\text{HTA}},{\text{db}} - {3}.{131}\;{R}^{{2}} = 0.{856}$$
(13)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = 0.{567 }\% {\text{HTA}},{\text{db}} - {1}.{9}0{1}\;{R}^{{2}} = 0.{81}0$$
(14)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = 0.{856}\;\% {\text{HTA}},{\text{db}} + {3}.{81}0\;{R}^{{2}} = 0.{681}$$
(15)

Pniowek DD samples (n = 14/15)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = 0.{214 }\% {\text{HTA}},{\text{db}} + {2}.{5}0{4}\;{R}^{{2}} = 0.{78}0$$
(16)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = 0.{2}.{83 }\% {\text{HTA}},{\text{db}} + {7}.{578}\;{R}^{{2}} = 0.{673}$$
(17)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = 0.{4}0{9 }\% {\text{HTA}},{\text{db}} + {3}0.{3}0{5} \;{R}^{{2}} = 0.{335}$$
(18)

Rest USCB DD samples (n = 23/28)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = 0.{23}0\% {\text{HTA}},{\text{db}} + {1}.{287}\;{R}^{{2}} = 0.{776}$$
(19)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = 0.{378 }\% {\text{HTA}},{\text{db}} + {4}.{196}\;R^{{2}} = 0.{728}$$
(20)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = 0.{494 }\% {\text{HTA}},{\text{db}} + {2}0.{191} \;R^{{2}} = 0.{433}$$
(21)

All USCB DD samples (n = 61/69)

$${\text{DD}}_{{{2}.{5}}} \left( {\% {\text{v}}} \right) = 0.{31}0 \, \% {\text{HTA}},{\text{db}} - 0.{61}0 \; {R}^{{2}} = 0.{759}$$
(22)
$${\text{DD}}_{{4}} \left( {\% {\text{v}}} \right) = 0.{431}\% {\text{HTA}},{\text{db}} + {1}.{529}\; {R}^{{2}} = 0.{733}$$
(23)
$${\text{DD}}_{{{1}0}} \left( {\% {\text{v}}} \right) = 0.{67}0 \, \% {\text{HTA}},{\text{db}} + {12}.0{7}\; {R}^{{2}} = 0.{5}00$$
(24)

A correlation was found between the proportions of finer fractions in DD500 and the ash yield for the USCB samples. An increase in the HTA yield was positively correlated with levels of the finer PSD, which is most obvious for the DD2.5 size fraction (8 samples of a total of 69 have not been considered in these calculations being outliers in data from each of the mines) (Eqs. (10)–(24)). This fact makes sense because the coal matrix (organic fraction) breaks to give coarser dust particles than clay, quartz and other mineral species, which increases the ash yields. Furthermore, the very low moisture contents of these DD500 samples allow finding these ash and high loads of fine DD fractions (Palmer et al. 1990; Jiang and Sheng 2018; Trechera et al. 2020). Another possible reason for the correlation between particle size and ash content in the USCB, not related to moisture content, could be that, given the high ranks of these coals, moisture is associated with the inorganic part, namely mineral matter (Jiang and Sheng 2018). However, this seems not to be the case here because of the lack of correlation between moisture and ash yields in most mines from the USCB, and the negative correlations found in two of them (Marcel and Bobrek-Piekary, Table 2). It is important also to highlight that when considering samples from all the mines from the USCB, the marked positive correlation is still evident (R2 = 0.76, 0.73 and 0.50 for DD2.5, DD4 and DD10, Eqs. (22)–(24)), suggesting a regional similarity across the entire USCB coal basin, independently of the 15 individual mines sampled.

A multilinear regression analysis was performed both between moisture and ash contents for each mine separately (in USCB), and by method of collection (in VCB). No correlations were observed showing that moisture content is not provided by the coal samples (with the exception of Marcel and Bobrek-Piekary mines, previously commented).

4.2 Geochemical characterisation

4.2.1 Velenje coal basin samples

The concentrations of different elements in the coal sample generally lie within the range for worldwide coals presented by Ketris and Yudovich (2009) (Table S9), with a few minor exceptions (Tables S10, S11 and S12) such as for P (higher by a factor of 1.2), Zn (1.4), U (1.6) and Mn (1.7).

4.2.2 Deposited dust samples (DD500)

Figure 3a and b show a comparison between major and trace element concentrations in coal dust and parent coal samples. Most of the major elements show higher (2 × or more) concentrations in the coal dust, with the exception of S which is very similar. This could relate to either: (i) a higher content in mineral matter, with elements such as Al, Na, P or K coming from the coal gangue; (ii) the mixing of coal dust with coal fly ash (Malta fly ash is commonly used in the mines as backfill, and Calplex, a carbonate ‘rock dust,’ is used for fire hazard reduction), that is chemically enriched in Al, Mg, Fe, Ca, Na, P, and K compared to coal (Table S13 in Supplementary Material); (iii) machine wear, especially for elements such as Fe and Mg.

Fig. 3
figure 3

Average deposited dust/parent coal elemental ratios for all dust fractions and standard deviation for the major (a) and trace (b) elements in the Velenje Coal Basin samples. Dashed red line represents ratio = 1. Average deposited dust/parent coal ratios from the different methods of collection (“fresh” plastic tray and “aged” mine surface deposits collected with a brush) for the major (c) and trace (d) elements in the Velenje Coal Basin. In black, average ratios from all deposited dust coal samples. In red, average ratios from all the “fresh” deposited dust. In green, average ratios from all the “aged” deposited dust

With regards to trace elements, again most of them are enriched in the coal dust compared with parent coal samples by 2 × or more, except for Mo and Sr. The trace elements Sn, Se, Ni, Zr and Sb show the highest relative enrichments (> 5x), that may be due to contamination from machinery (belts, extractors, drills etc.) in the underground mine. Another potential contaminant is again fly ash (Malta), especially in trace elements such as Cr, Y, Ni, Se and Zr (Table S14).

Figure 3c and d compare the trace element contents of DD500 versus those in the parent coal for samples obtained using different collection methods: on plastic trays (fresh) or for longer term (aged) samples deposited onto mine surfaces and collected using a brush. Major elements, and the majority of the trace elements, do not show any appreciable difference according to the coal dust collection method. In the exceptional case of Sb (Fig. 3d), one might speculate that this could relate to the brushing collection method which may have picked up more brake-contaminated dust on the floor of the mine than that obtained from the collection plastic tray placed 1.5 m above ground level.

Finally, there are large variations in the elemental concentrations of some DD500 samples. This could be related to the types of mining activities carried out in the specific areas where sampling was carried out. Samples collected from the longwall k.-95/A-conveyor transfer point and longwall k.-95/B-conveyor transfer point present lower DD500 concentrations compared with DD samples from other locations. The sample ROCD_PV_002_4, collected from the longwall k.-95/A-conveyor transfer point, has the lowest element concentrations. This is probably because, in these zones, mining activity is less intense, and for this reason the levels of most elements (except Sb and U) decrease.

4.2.3 Respirable deposited dust (RDD4 and RDD2.5) samples

The successful separation of the respirable fractions from the deposited coal dust (RDD4 and RDD2.5), using our particle size separator device, is demonstrated by the size distributions plotted on Figure S1 and shown in Table S15 in Supplementary Material. Concentrations of major and trace elements (Tables S13 and S14 in Supplementary Material) in these respirable (RDD4 and RDD2.5) samples, and the specific RDD4/DD500 ratios are shown in Fig. 4a, where major and trace elements (especially Ba, Se, Zn, Pb, As, Sn, Cu and Sb) are shown to be typically enriched in the finer dust fraction (ratios RDD4/DD500 1.2–1.4: Fig. 4a).

Fig. 4
figure 4

a Average RDD4/DD500 major and trace element ratios and standard deviation for the Velenje Coal Basin samples. b Average RDD2.5/RDD4 major and trace element ratios and standard deviation for the Velenje Coal Basin samples

Further fractionation into the RDD2.5 size fraction is less apparent for the major elements (Fig. 4b, top), but for the trace elements (Fig. 4b, bottom) there are further enrichments in the metallic and metalloid elements, especially Ni, Sn and Cu. Common patterns were found concerning the elements enriched in the RDD fractions compared to the DD500. This is mainly the case for Zn, Sb, As, Sn and Pb which were probably mostly sourced from the wear of mining machinery and from sulphide minerals in the coal. In contrast, Mn is the only trace element which decreases in its concentration in the finest fractions (RDD4 and RDD2.5) when compared to DD500 (Fig. 4a and 4b, bottom). This result is similar to previous results from Trechera et al. (2020, 2021). Mn is probably joined with organic matter (producing coarser particles), which is commonly associated with carbonate minerals (Swaine 1990).

4.2.4 Upper silesian coal basin samples

Representative coal samples from five of the USCB mines (Pniowek, Jankowice, Marcel, Bielszowice and Murcki-Staszic) were analysed (Tables S10, S11 and S12). Elemental concentrations were generally < 1.5 × the average worldwide coal concentrations reported by Ketris and Yudovich (2009) (Table S9), although only Bielszowice mine samples meet the < 1.5 factor for all elements. Other exceptions are for Mn (1.5 × in the Marcel mine), Zn, Pb and Co (1.6x, 1.8x, and 2.8 × respectively in Jankowice mine), Y (1.8x), Zn (2.1x), Sr (2.6x), Ba (2.8x), Sn (3.2x) and P (6.7x) in the Pniowek mine, and Ba (1.5x), Co (1.5–1.7x), W (1.6x), Cu (2.0x), Sb (2.2x), Pb (1.8–4.9x), Zn (6.4x) and Mn (5.0-12x) in Murcki-Staszic.

4.2.5 Deposited dust samples (DD500)

When DD500 samples (Table S16 and S17) are compared with their parent coals, most major elements in all USCB DD500 samples had increased concentrations, by at least two times, when compared with parent coals, with the exception of S, and P (in Pniowek), which reached 0.4–1.4 times, and 0.9 times, respectively.

In the case of S, its content in coal dusts was lower than (or similar) to the parent coal in all USCB mines. This is likely related to S mostly occurring in the coal matrix and therefore coming from the organic part of the DD500, and accordingly these (enriched in mineral matter) contain lower S than the parent coals. The content in pyrite in these coal dusts is presumably minimal due to their low S loads (Kotarba 2001). Ca shows a strongly increased concentration in the DD500 samples when compared to the parent coals (20–300 times). This effect is attributed to the walls of these USCB mines being shotcrete with concrete to prevent collapses, although additional input of calcium carbonate from the coal gangue cannot be excluded. In addition to this, carbonate ‘rock dust’ is spread in some parts of the underground mines to prevent explosions. Also K concentrations are markedly higher (4 times) in the DD500 compared with their parent coals in all USCB underground mines, with the exception of the Jankowice mine. This observation is attributed to the higher mineral matter loads in the dust.

With regard to trace element concentrations, most trace element concentrations were higher or very similar in DD500 than in the parent coal, with the exceptions of As, Co, Pb, Mn and Ge in the Murcki-Staszic mine; Co and Ni in the Marcel mine and As, Co, Mo and Ni in the Jankowice mine.

4.2.6 Respirable deposited dust samples (RDD4)

Average concentrations of major elements in RDD4 and DD500 samples again show a marked enrichment in Al, Ca, Na, K, P in the RDD4 fraction (Tables S16 to S19), with the exception of S, Fe and Mg which are enriched in the DD500, and accordingly in the coarser fractions (Fig. 5a). Thus, in 13 (S), 11 (Fe) and 10 (Mg) out of 15 USCB mines, RDD4/DD500 ratios are < 1 (Fig. 5b). In contrast, Al ratio generally slightly exceeds 1, and this ratio is even > 2 in some cases for Ca, Na, K and P.

Fig. 5
figure 5

a Averaged major element RDD4/DD500 ratios for all samples from the Upper Silesian Coal Basin. b Averaged major element RDD4/DD500 ratios for samples from individual mines in the Upper Silesian Coal Basin. c Averaged trace element RDD4/DD500 ratios for all samples from Upper Silesian Coal Basin. d Averaged trace element RDD4/DD500 ratios for samples from individual mines in the Upper Silesian Coal Basin. PZ, Piast-Ziemowit; KN, Knurów; ST, Murcki-Staszic; M, Marcel; Z, Zofiowka; BP, Bobrek-Piekary; BL, Bielszowice; POK, Pokój; BUD, Budryk; CHW, Chwalowice; PN, Pniowek; JN, Jankowice; JAS, Jastrzebie; HAL, Halemba; MW, Myslowice-Wesola

RDD4/DD500 ratios for trace element concentrations are shown in Fig. 5c and d, with considerable variations, but a clear trend towards the enrichment of metals in the RDD4. Elements such as Cu, Sn, Zn and As are concentrated by 1.5–1.7 times, but it is the metalloid Sb that is most enriched (2.5 times). However, Fig. 5c also highlights major geochemical variations (high standard deviations) for average dusts. The least fractionated metallic element is Mn, with enrichments in only five mines (Marcel, Chwalowice, Pniowek, Jankowice, and Jastrzebie). In contrast, Zr, Se and Ti reach RDD4/DD500 ratios > 2 in some cases; and Cu, Sn, Zn, As and Sb often reach > 3.

In summary, the results show that neither the parent coal, nor the DD500 have geochemical patterns reproducing those of the respirable dust (RDD4), specially for some metals.

4.3 PM10 measurements

4.3.1 Velenje coal basin PM10 concentrations

PM10 airborne concentrations in the VCB mine were measured during coal working using CIP10 samplers at different distances from long-wall shearer operations (50, 100, 150 and 200 m). Here, the shearer moves along to cut a section of the coal seam then is progressed forwards by large hydraulic push rams, attached to roof supports, to cut the next section (Reid et al. 2006).

No reference on occupational exposure limits was found for ambient concentrations of metals in Slovenia. According to the occupational exposure limits for airborne toxic substances from Poland (Basic Legal Act in Poland 2010), measured elemental concentrations did not exceed the respective limit values.

As was expected, ambient PM10 concentrations progressively decreased with increasing distance from the long-wall shearer operations (Table 1). The general tendency was for constant element concentrations in PM10 in the VCB mine at a distance of 150 m from the shearer. At a distance of 200 m there was a slight decrease in their concentrations which is likely to have been due to the effectiveness of the long-wall "U-type" ventilation system, with a flow of new air coming into the workings (McPherson 1993; Smith et al. 1994; Marts et al. 2015), as well as the shearer spraying system reducing airborne dust concentrations (Colinet et al. 2010). There were a few exceptions to this such as for Li and Zn (50 m away) and Cr, Ni and Cu (150 m away) where there were large increases in concentrations, with no clear explanation.

Table 1 PM10 levels and major element oxide and trace element concentrations in PM10 per m3 of air in the Velenje Coal Basin mine during shearer-longwall operations at a T-junction (exhaust air from longwall face), longwall k-95

Activity distance (m)

0

50

100

150

200

PM10 (µg/m3)

5410

5135

4814

4994

3500

Sampling minutes

105

100

95

83

155

Coal (µg/m3)

1400

1317

1280

1242

878

SiO2 (µg/m3)

167

167

174

169

119

Al2O3 (µg/m3)

112

111

116

113

79

CaO (µg/m3)

100

88

97

97

62

Fe2O3 (µg/m3)

132

125

134

123

95

K2O (µg/m3)

94

67

13

13

8.4

MgO (µg/m3)

18

16

18

17

11

Na2O (µg/m3)

31

37

31

38

25

S (µg/m3)

120

117

124

115

85

TiO2 (µg/m3)

4.7

4.4

5.1

4.3

3.1

Li (ng/m3)

35

49

29

36

26

Sc (ng/m3)

 < 2

 < 2

 < 2

 < 2

 < 2

V (ng/m3)

85

79

84

80

59

Cr (ng/m3)

132

125

145

222

95

Mn (ng/m3)

839

780

800

745

553

Co (ng/m3)

11

9.9

10

9.9

8.0

Ni (ng/m3)

20

14

18

127

18

Cu (ng/m3)

124

156

171

245

140

Zn (ng/m3)

1026

1871

1009

1042

902

Ga (ng/m3)

18

17

18

18

13

As (ng/m3)

35

41

38

41

29

Rb (ng/m3)

91

88

88

83

65

Sr (ng/m3)

443

379

377

441

280

Y (ng/m3)

15

14

15

14

9.8

Zr (ng/m3)

76

50

51

46

42

Nb (ng/m3)

10

 < 2

 < 2

 < 2

6.5

Mo (ng/m3)

74

70

73

66

56

Sb (ng/m3)

636

655

642

578

621

Cs (ng/m3)

 < 2

 < 2

 < 2

 < 2

6.8

Ba (ng/m3)

296

286

330

293

221

Pb (ng/m3)

83

72

53

76

52

Th (ng/m3)

 < 2

 < 2

 < 2

 < 2

 < 2

U (ng/m3)

28

26

29

25

19

REE (ng/m3)

90

89

94

91

68

Another interesting pattern is the higher ambient concentrations of Mn and Fe2O3 in comparison with similar studies carried out in Chinese underground coal mines (Trechera et al. 2020, 2021), which are possibly from machine wear or from higher Mn and Fe contents in the Slovenian coals.

4.3.2 Upper silesian coal basin PM10 concentrations

Ambient PM10 concentrations in the different USCB underground coal mines were measured during two different coal mining activities: i) Shearer cutting during long-wall excavations (as for the VCB mine), ii) Roadway drivage, which consists of cutting the coal and rock using a road-header, loading and transportation of the excavated material and the installing of underground supports.

Tables 2, 3, 4 show PM10 concentrations in the Knurów, Murcki-Staszic, Pnioweck, Marcel, and Bielszowice underground coal mines. Ambient concentrations of all determined elements are lower than the Polish recommended maximum values for occupational environments (Basic Legal Act in Poland 2010). However, ambient PM10 concentrations should not exceed 4.0 mg/m3 where the coal contains 2%–10% of SiO2, or 2.0 mg/m3 with 10%–50% of SiO2 over a period of 8 h. PM10 concentrations measured in various of the studied mines were above these values, such as in the Pniowek mine (9.8 to 12.0 mg/m3 coal dust concentration in zone b, 4.7 to 6.0 mg/m3 in zone a and 14 to 24 mg/m3 in zone c, Table 3), Marcel mine (8.3 to 9.6 mg/m3 in zone a, and 12 to 13 mg/m3 in zone b), Knurów mine (7.7 to 14.0 mg/m3, Table 2),) and Bielszowice mine (3.3 to 6.1 mg/m3, Tables 3 and 4). However, these concentrations were measured over periods shorter than 8 h, and therefore are not comparable with the occupational standard. In any case, these short time high concentrations reinforced the importance of the protocols applied by mining companies to protect coal workers with sophisticated personal protective equipment and coal dust abatement controls.

Table 2 PM10 levels and major element oxide and trace element concentrations during roadway drivage technique in Knurów mine (exhaust air from roadway face; 5b, seam 404/1) and shearer cutting- longwall advance in Murcki-Staszic mine at the T-junction (exhaust air from longwall face 03/seam 510) from the Upper Silesian Coal Basin

Mine

Knurów

Knurów

Knurów

Murcki-Staszic

Murcki-Staszic

Murcki-Staszic

Murcki-Staszic

Murcki-Staszic

Activity distance (m)

20–50

100–140

220–280

0

50

190

280

350

PM10 (µg/m3)

14,025

15,000

7702

11,905

7833

14,969

6578

7963

Sampling minutes

80

80

95

95

87

97

94

99

Coal (µg/m3)

6221

6086

3299

613

538

843

672

486

SiO2 (µg/m3)

2669

2404

927

151

117

228

125

106

Al2O3 (µg/m3)

1780

1603

618

101

78

152

83

71

CaO (µg/m3)

296

751

1128

272

217

331

388

214

Fe2O3 (µg/m3)

712

651

272

74

79

106

50

66

K2O (µg/m3)

273

250

93

5.6

8.1

13

10

13

MgO (µg/m3)

180

151

69

32

32

44

27

22

Na2O (µg/m3)

146

121

96

57

55

86

61

52

S (µg/m3)

49

48

53

47

61

54

54

41

TiO2 (µg/m3)

89

82

32

7.1

5.5

9.7

8.1

5.1

Li (ng/m3)

575

566

206

99

74

91

50

79

Sc (ng/m3)

168

150

50

 < 2

 < 2

 < 2

 < 2

 < 2

V (ng/m3)

1048

979

345

58

40

83

46

37

Cr (ng/m3)

1243

1130

426

75

89

654

80

78

Mn (ng/m3)

10,646

9736

3593

511

467

748

421

455

Co (ng/m3)

176

156

55

 < 2

 < 2

17

9.7

6.9

Ni (ng/m3)

446

375

152

 < 2

6.87

553

 < 2

 < 2

Cu (ng/m3)

398

547

374

 < 2

848

93

81

31

Zn (ng/m3)

1569

1419

1020

488

1611

2179

4686

1366

Ga (ng/m3)

242

225

75

 < 2

12

21

13

10

As (ng/m3)

51

46

25

 < 2

2.2

143

 < 2

 < 2

Rb (ng/m3)

1206

1086

384

36

33

62

37

29

Sr (ng/m3)

1248

1532

1277

509

1021

764

589

426

Y (ng/m3)

243

228

83

 < 2

 < 2

 < 2

11

10

Zr (ng/m3)

1111

871

296

38

44

64

43

32

Nb (ng/m3)

137

122

44

 < 2

 < 2

 < 2

 < 2

 < 2

Mo (ng/m3)

 < 2

 < 2

 < 2

 < 2

 < 2

 < 2

12

 < 2

Sb (ng/m3)

133

245

246

626

965

897

393

738

Cs (ng/m3)

73

65

25

 < 2

 < 2

 < 2

 < 2

 < 2

Ba (ng/m3)

4368

3949

1555

806

829

1358

9778

1089

Pb (ng/m3)

277

247

142

10

70

157

113

46

Th (ng/m3)

124

121

42

 < 2

 < 2

 < 2

 < 2

 < 2

U (ng/m3)

41

38

16

 < 2

 < 2

 < 2

 < 2

 < 2

REE (ng/m3)

1724

1562

588

68

67

130

71

66

Table 3 PM10 levels and major element oxide and trace element concentrations in the Pnioweck mine from the Upper Silesian Coal Basin. Zone a) Roadway drivage technique (exhaust air from roadway face PW-1, seam 358/16). Zone b) Shearer cutting—longwall advance at the T-junction (exhaust air from longwall face B-4, seam 404/2). Zone c) Roadway drivage technique (exhaust air from roadway face seam K-9, seam 363 362/3)

Zone

a

a

a

b

b

b

b

c

c

c

c

c

Activity distance (m)

60

120

180

0

50

100

150

0

50

100

150

200

PM10 (µg/m3)

4669

4602

5975

10,544

11,464

10,325

9843

24,343

10,474

17,480

17,710

13,405

Sampling minutes

96

93

85

125

125

120

115

105

103

70

60

83

Coal (µg/m3)

1794

1721

2193

2346

2136

2975

3145

9245

5919

7052

7589

6329

SiO2 (µg/m3)

809

783

1013

566

534

834

861

1969

1247

1493

1603

1317

Al2O3 (µg/m3)

539

522

675

377

356

556

574

1312

831

995

1068

878

CaO (µg/m3)

194

154

168

230

197

156

187

243

164

182

179

177

Fe2O3 (µg/m3)

184

186

250

68

70

111

106

353

212

266

294

239

K2O (µg/m3)

79

75

99

84

44

68

72

179

109

138

148

119

MgO (µg/m3)

41

35

56

18

17

19

27

81

49

51

62

57

Na2O (µg/m3)

50

56

62

38

40

30

48

71

66

74

89

77

S (µg/m3)

44

49

46

49

44

41

41

98

80

88

98

86

TiO2 (µg/m3)

27

27

34

25

22

29

34

61

37

44

46

40

Li (ng/m3)

264

266

350

238

221

335

318

859

607

695

699

672

Sc (ng/m3)

48

47

61

38

37

53

52

136

79

102

104

82

V (ng/m3)

405

410

509

572

525

554

559

1307

827

974

995

822

Cr (ng/m3)

372

351

597

436

359

414

441

957

605

710

770

619

Mn (ng/m3)

2012

2098

3036

483

471

780

752

3654

1997

2555

2827

2262

Co (ng/m3)

57

55

74

92

89

75

80

169

107

126

136

111

Ni (ng/m3)

235

69

316

288

172

205

197

542

348

382

416

305

Cu (ng/m3)

125

10

1104

178

70

115

149

295

178

176

82

101

Zn (ng/m3)

1187

828

1389

1136

321

268

706

2156

1801

2744

1845

2217

Ga (ng/m3)

77

80

98

64

62

85

86

198

122

147

155

129

As (ng/m3)

40

17

84

72

7.5

10

34

311

110

154

141

84

Rb (ng/m3)

403

391

514

275

253

415

405

1066

635

779

818

677

Sr (ng/m3)

660

597

726

718

694

683

754

1798

1223

1416

1509

1277

Y (ng/m3)

73

70

96

53

54

69

73

204

130

149

161

130

Zr (ng/m3)

373

343

493

188

189

265

489

816

479

547

591

480

Nb (ng/m3)

46

44

57

45

42

53

65

121

72

82

88

72

Mo (ng/m3)

 < 2

 < 2

 < 2

 < 2

 < 2

 < 2

 < 2

32

9

 < 2

 < 2

 < 2

Sb (ng/m3)

296

265

305

340

358

258

318

888

634

606

584

682

Cs (ng/m3)

41

42

51

34

31

48

45

104

67

79

86

69

Ba (ng/m3)

1816

1765

2354

1977

1961

2196

2243

5695

3489

4064

4239

3590

Pb (ng/m3)

116

148

291

240

109

99

145

232

140

141

184

125

Th (ng/m3)

43

42

46

25

27

36

56

101

64

71

80

65

U (ng/m3)

15

 < 2

19

 < 2

 < 2

 < 2

 < 2

42

27

32

35

13

REE (ng/m3)

609

562

799

415

404

576

586

1537

941

1130

1131

959

Table 4 PM10 levels and major element oxide and trace element concentrations in different zones of Marcel (M) and Bielszowice (BL) mines from the Upper Silesian Coal Basin. M zone a: Shearer cutting—longwall advance at the T-junction (exhaust air from longwall face W-7, seam 505); M zone b): Roadway drivage technique (exhaust air from roadway face C3, seam 507); and BL: Roadway drivage technique (exhaust air from roadway face 4az, seam 510wd)

Mine-zone

M-a

M-a

M-a

M-b

M-b

M-b

M-b

BL

BL

BL

Location

20 m

70 m

120 m

10 m

40 m

60 m

80 m

15 m

40 m

60 m

Concentration (µg/m3)

9590

8300

8255

12,181

12,960

12,505

12,242

6116

5627

3320

Sampling minutes

160

150

140

105

100

95

95

155

150

140

Coal (µg/m3)

2018

1788

1814

4514

4812

4347

4678

1757

3492

1214

SiO2 (µg/m3)

450

396

398

1502

1625

1440

1548

425

853

811

Al2O3 (µg/m3)

300

264

265

1001

1083

960

1032

283

569

541

CaO (µg/m3)

153

128

124

221

230

226

264

132

213

192

Fe2O3 (µg/m3)

105

100

107

344

339

325

337

65

134

128

K2O (µg/m3)

29

30

28

111

114

103

110

28

52

50

MgO (µg/m3)

30

26

25

74

80

72

79

11

34

31

Na2O (µg/m3)

59

53

58

24

26

31

29

6.1

17

13

S (µg/m3)

64

56

64

20

20

20

21

 < 2

20

21

TiO2 (µg/m3)

13

12

12

42

45

40

43

16

29

28

Li (ng/m3)

91

81

64

495

420

310

399

119

187

156

Sc (ng/m3)

24

19

20

87

87

80

86

22

45

39

V (ng/m3)

196

177

176

648

670

593

642

194

302

285

Cr (ng/m3)

229

225

214

625

684

595

622

97

306

277

Mn (ng/m3)

951

927

946

2548

2544

2373

2544

598

1154

1082

Co (ng/m3)

31

27

26

82

79

74

80

59

126

118

Ni (ng/m3)

132

130

222

263

259

245

247

11

150

114

Cu (ng/m3)

96

640

108

572

629

573

520

 < 2

147

148

Zn (ng/m3)

4584

4817

4600

3482

3497

3211

3625

478

1248

1101

Ga (ng/m3)

43

39

41

128

144

120

130

40

76

70

As (ng/m3)

33

100

52

34

48.7

39

38

25

161

62

Rb (ng/m3)

171

150

151

569

593

525

570

117

212

202

Sr (ng/m3)

675

550

534

896

941

896

911

346

602

548

Y (ng/m3)

46

41

43

113

125

110

119

33

53

49

Zr (ng/m3)

142

139

128

499

533

463

503

118

252

248

Nb (ng/m3)

22

21

20

65

71

62

67

22

42

39

Mo (ng/m3)

17

21

17

 < 2

 < 2

 < 2

 < 2

 < 2

 < 2

 < 2

Sb (ng/m3)

2292

2030

2361

1539

1311

1279

1467

204

616

611

Cs (ng/m3)

24

21

21

71

75

65

71

 < 2

15

14

Ba (ng/m3)

1477

1302

1301

4011

4037

4022

3458

917

1400

1320

Pb (ng/m3)

67

164

57

114

132

97

104

7.3

85

86

Th (ng/m3)

21

21

18

50

58

53

55

24

38

37

U (ng/m3)

16

14

14

32

33

30

32

 < 2

17

15

REE (ng/m3)

310

269

270

897

962

861

912

371

625

592

As expected, PM10 concentrations are mostly higher closer to mine machine operations (Tables 2, 3, 4), with the exception of the Pniowek mine zone a, where measurements took place during a maintenance shift when mine machinery was switched off but the ventilation system continued to operate.

Comparing the two activities studied in all mines, shearer cutting during long-wall excavations emitted less PM10 coal dust than the roadway drivage operations. Mn concentrations were also clearly higher during roadway drivage. When comparing the different coal mines, lower PM10 concentrations were recorded in three mines, Murcki-Staszic, Pniowek zone b and Bielszowice. In contrast, in the Knurów mine, concentrations of Al2O3, CaO, Fe2O3, K2O, MgO, Na2O, as well as metals such as Pb and Cr, were higher in comparison with other mines from this and other studies (Trechera et al. 2020, 2021).

When PM10 levels produced by each mining activity are compared, elevated levels of As (2.2 times), Ba (1.5 times), and Li (2.2 times) are evidenced in the Pniowek zone c, and Zn (1.7–2.3 times, Ba (1.5 times) and Cu (2 times) in the Marcel mine (zones a and b). The elevated levels of these metals were likely derived from a mix of sources including machine wear (Cu, Zn, Ba), resuspension of DD from the floor and walls (As, Ba, Li), and PM brought from other zones of the mine by the ventilation system.

4.4 PM10 and RDD4 comparison

The aim of comparing PM10 and RDD4 concentrations was to find the possible sources of each particle size fraction to clarify the dominant type of particle present in underground coal mine air, always in a qualitative way. The comparison was performed by calculating the difference in PM10 concentrations (Tables 1, 2, 3, 4) of major and trace elements.

In the VCB underground mine, all element concentrations in PM10 were lower than in the RDD4. Some elements, such as K, Cu, Zn, and Mo, were similar in both PM10 and RDD4 samples, whereas Na and Sb show higher concentrations in PM10 from all sampling locations (1.3–2.0 times Na and 7.1–12 times higher Sb).

In the case of the USCB mines, most elements in the Knurów mine show lower concentrations in the PM10 samples compared to RDD4, with the exception of Ti, V, Mn, Cr, Co, Mg, Ca and REE in some specific locations. The Murcki-Staszic mine also showed most elements to be enriched in the RDD4 samples, except for Sb, Na, and S with higher concentrations in the PM10 samples in all locations (1.7 to 4.0 times for Sb, 1.4 to 2.8 times for Na and 1.7 to 4.0 times for S). PM10/RDD4 ratios in the Pniowek mine zone were normally between 1.0 to 2.0, with the exception of Ba, Sr, Ca, Sb and Co that had a ratio < 1.0, and S, Mn and Na with ratios > 2.0 (1.7 to 2.4, 2.3 to 2.8, 3.3 to 3.9, respectively). In zone b, all PM10/RDD4 ratios were below or close to 1 and in zone c Ni, Cu, Zn, As, Sr Mo, Pb, U and Ca concentrations were lower in the PM10 than RDD4.

In the Marcel mine zone a, only Zn (3.9 to 4.7), Sb (6.9 to 8.2) and Na (4.4 to 5.0) were markedly higher in the PM10 than RDD4, whereas Li, V, Cr, Mn, Cu, As, Sr, Y, Mo and Pb were enriched in the RDD4 samples. In contrast, in Marcel mine zone b, a large number of elements such as Li, Sc, Ti, Ga Rb, Zr, Nb, Cs, Ba, Al, K and Sb, show PM10/RDD4 ratios higher than 2.5, especially in the case of Sb which had ratios of 3.5 to 4.4. Finally, in Bielszowice, most elements showed higher concentrations in the PM10 samples compared to their respective RDD4 sample, with the exception of Ca, Mg, Na, Sr and Pb.

In summary, some elements such as Sb, Zn, Mn, Na, and S showed elevated concentrations in PM10 samples compared with their respective RDD4. Firstly, maybe all particle sizes of these elements are coarser, for this reason their concentrations would be higher in PM10. Moreover, the presence of airborne Sb, Zn and Mn could be directly related to wear in machinery used in the underground mines. Amato et al. 2009a, b, among others, show Sb, Mn and Zn in atmospheric PM from abrasion sources (i.e. brake pads or tyres in vehicles) have generally a coarser size (PM2.5–10) than resuspension. If this is the case in our studies (machinery wear sources), higher PM10 than DD4 levels might be expected for them. On the other hand, Na and S are directly related to coal handling activities because these elements are coming from their organic matter. Furthermore, a multi regression analysis was performed to correlate average S content in PM10 with S in parent coal from each coal mine. The high correlation (R2 = 0.98, P = 0.0001) indicates that S emissions related to PM10 ambient concentrations in the underground coal mine are directly related to coal handling and manipulation (drilling, extraction, transportation).

5.Conclusions

The analysis of the particle size distribution (PSD) of deposited dust (DD) finer than 500 µm (DD500) in underground mines confirms the need to wet airborne coal dust and DD. In the Velenje Coal Basin (VCB) underground mine, DD500 is wetted which reduces the load of respirable DD (RDD), and accordingly the potential for miners to inhale coal dusts and for coal dust explosions. In contrast, in the Upper Silesian Coal Basin (USCB) underground mines, the combination of low moisture in the DD500 and a high ash yield doubles the relative load of RDD in DD500 when compared with VCB.

The geochemical composition of aged (collected over long time periods onto existing surfaces; sampled with a brush) and fresh (deposited on PVC trays over short time periods) DD coal samples from the same locations did not show major differences with the exception of Sb which was enriched in the aged DD, probably due to a relatively higher component of dust from brake-wear.

A comparison of geochemical compositions of parent coals and DD500 revealed very important differences, with relatively high levels of most major and trace elements in the DD500 for the VCB underground mine, especially for Ni, Zr and Sb. This can be attributed to an increased contribution to DD500 from particulate matter (PM) emitted from the wear of mine machinery (e.g. drills, excavators and blts, coal gangue, and from coal fly ash (called Malta D) and carbonate ‘rock dust’ (called Calplex) commonly used as backfill and fire hazard reduction respectively in this mine. The chemical composition of RDD (deposited dust finer than 4 and 2.5 µm, RDD4 and RDD2.5) samples extracted in the laboratory from DD500 reveals frequent enrichments of metals and metalloids in the RDD. In the VCB mine these include Ba, Se, Zn, Pb, As, Cr, Mo, and especially Ni, Sn, Cu, and Sb.

Similar results were obtained for the USCB mines, where DD500 samples were also markedly enriched in Ca and K compared with parent coals. The enrichment in Ca is attributed to dusts from concrete gunite used behind mine supports, whereas K is attributed to dust contributions from coal gangue. An enrichment in Rb, Zr, Sb, Zn, Cr, V, Ti, Mn and Sn in the DD500 compared with the parent coal was evidenced. In contrast, S contents were higher in the parent coals probably due to the major association of S with the coal matrix (reduced in DD with respect the coal seam).

The USCB mines also showed a marked enrichment in metals and metalloids in RDD4 compared with DD500, especially for Cu, Sn, Zn, As and Sb (> 1.5 fold higher than in the DD500). In contrast, Mn was reduced in the RDD4 samples compared to DD500 in all samples from both VCB and USCB mines.

According to the above results, we conclude that the geochemical composition of the RDD is markedly different to that of DD and the parent coals, and that the method we have used allows extracting RDD from DD sampling for further PSD, chemical and mineralogical analyses.

Airborne PM10 concentrations were below established occupational exposure limits. Long-wall shearer excavations produce lower PM coal dust concentrations than roadway drivage operations. Higher PM10 concentrations were observed in some USCB underground mine sections related to different mechanical tasks being conducted in the mines. However, these concentration peaks were present only transiently (typically over 80–135 min) and thus are not directly applicable to the 8 h period legislated for occupation exposure.

Concentrations of metals and metalloids in PM10 were normally lower than in the RDD4 samples, with the exception of Sb, Zn, Mn, Na, and S in some mines, that they could be present in the coarser particles. The presence of Sb, Zn, and Mn in PM10 may be attributed to machine wear. In contrast, S and Na may be attributed to the organic matter of coal. Moreover, S concentrations are directly correlated with coal handling manipulation in the underground coal mines (R2 = 0.98, P < 0.05).

There have been few studies regarding the possible health implications of inhaling some of these metals and metalloids, further studies on the oxidative potential and cellular toxicology of coal dust samples should be carried out (Birch and Scollen 2003; Moreno et al. 2008; Colinet et al. 2010; Riley et al. 2012; Perret et al. 2017). Concentrations of such metals, and of coal dusts in general, should be kept at a minimum when working in underground mines. The studies carried out within the ROCD project aim to be the seed of future activities necessary to achieve better working conditions in this environment, especially given that global coal production may grow over the next few decades (BP 2020).

References

[1] Ajrash MJ, Zanganeh J, Moghtaderi B (2017) The effects of coal dust concentrations and particle sizes on the minimum auto-ignition temperature of a coal dust cloud. Fire Mater 41:908–915. https://doi.org/10.1002/fam.2437
[2] Amato F, Pandolfi M, Escrig A et al (2009a) Quantifying road dust resuspension in urban environment by Multilinear Engine: a comparison with PMF2. Atmos Environ 43:2770–2780. https://doi.org/10.1016/j.atmosenv.2009.02.039
[3] Amato F, Querol X, Alastuey A et al (2009b) Evaluating urban PM10 pollution benefit induced by street cleaning activities. Atmos Environ 43:4472–4480. https://doi.org/10.1016/j.atmosenv.2009.06.037
[4] Azam S, Mishra DP, Wang X et al (2019) Synergistic effect of surfactant compounding on improving dust suppression in a coal mine in Erdos, China. Process Saf Environ Prot 344:35–43. https://doi.org/10.1016/j.powtec.2018.12.061
[5] Basic Legal Act in Poland (2010) OCCUPATIONAL EXPOSURE LIMITS FOR AIRBORNE TOXIC SUBSTANCES. Basic Legal Act in Poland.
[6] Birch GF, Scollen A (2003) Heavy metals in road dust, gully pots and parkland soils in a highly urbanised sub-catchment of Port Jackson, Australia. Aust J Soil Res 41:1329–1342. https://doi.org/10.1071/SR02147
[7] BP (2020) bp Statistical Review of World Energy 2020
[8] Brezigar A (1987) Geologic setting of the Pre-Pliocene basement of the Velenje depression and its surroundings. Geologija 30:31–65
[9] Caballero-Gallardo K, Olivero-Verbel J (2016) Mice housed on coal dust-contaminated sand: a model to evaluate the impacts of coal mining on health. Toxicol Appl Pharmacol 294:11–20. https://doi.org/10.1016/j.taap.2016.01.009
[10] Cohen RAC, Patel A, Green FHY (2008) Lung disease caused by exposure to coal mine and silica dust. Semin Respir Crit Care Med 29:651–661
[11] Colinet JF, James PR, Jeffrey ML et al (2010) Best practices for dust control in coal mining. Centers Dis Control Prev Natl Inst Occup Saf Heal 01:17–36
[12] Dalal NARS, Newman J, Pack D et al (1995) Original Contrib 18:11–20
[13] EPA (1995) Reducing methane emissions from coal mines in poland: a handbook for expanding. Atmos Pollut Prev Div US Environ Prot AGENCY
[14] Fan L, Liu S (2021) Respirable nano-particulate generations and their pathogenesis in mining workplaces: a review. Int J Coal Sci Technol. https://doi.org/10.1007/s40789-021-00412-w
[15] Fan T, Zhou G, Wang J (2018) Preparation and characterization of a wetting-agglomeration-based hybrid coal dust suppressant. Process Saf Environ Prot 113:282–291. https://doi.org/10.1016/j.psep.2017.10.023
[16] Fang X, Yuan L, Jiang B et al (2020) Effect of water–fog particle size on dust fall efficiency of mechanized excavation face in coal mines. J Clean Prod 254:120146. https://doi.org/10.1016/j.jclepro.2020.120146
[17] Hu S, Gao Y, Feng G et al (2021) Experimental study of the dust-removal performance of a wet scrubber. Int J Coal Sci Technol 8:228–239. https://doi.org/10.1007/s40789-021-00410-y
[18] Jiang H, Luo Y (2021) Development of a roof bolter drilling control process to reduce the generation of respirable dust. Int J Coal Sci Technol 8:199–204. https://doi.org/10.1007/s40789-021-00413-9
[19] Jiang L, Sheng C (2018) Correlation of the sub-micrometer ash yield from pulverized coal combustion with coal ash composition. Energy Fuels 32:9961–9970. https://doi.org/10.1021/acs.energyfuels.8b02098
[20] Jin L, Liu J, Guo J et al (2021) Physicochemical factors affecting the wettability of copper mine blasting dust. Int J Coal Sci Technol 8:265–273. https://doi.org/10.1007/s40789-021-00411-x
[21] Johann-Essex V, Keles C, Rezaee M et al (2017) Respirable coal mine dust characteristics in samples collected in central and northern Appalachia. Int J Coal Geol 182:85–93. https://doi.org/10.1016/j.coal.2017.09.010
[22] Jonek-Kowalska I (2018) How do turbulent sectoral conditions sector influence the value of coal mining enterprises? Perspectives from the Central-Eastern Europe coal mining industry. Resour Policy 55:103–112. https://doi.org/10.1016/j.resourpol.2017.11.003
[23] Kedzior S (2015) Methane contents and coal-rank variability in the Upper Silesian Coal Basin, Poland. Int J Coal Geol 139:152–164. https://doi.org/10.1016/j.coal.2014.09.009
[24] Kędzior S (2009) Accumulation of coal-bed methane in the south-west part of the Upper Silesian Coal Basin (southern Poland). Int J Coal Geol 80:20–34. https://doi.org/10.1016/j.coal.2009.08.003
[25] Kędzior S (2019) Distribution of methane contents and coal rank in the profiles of deep boreholes in the Upper Silesian Coal Basin, Poland. Int J Coal Geol 202:190–208. https://doi.org/10.1016/j.coal.2018.12.010
[26] Ketris MP, Yudovich YE (2009) Estimations of clarkes for carbonaceous biolithes: world averages for trace element contents in black shales and coals. Int J Coal Geol 78:135–148. https://doi.org/10.1016/j.coal.2009.01.002
[27] Kotarba MJ (2001) Composition and origin of coalbed gases in the Upper Silesian and Lublin basins, Poland. Org Geochem 32:163–180. https://doi.org/10.1016/S0146-6380(00)00134-0
[28] Kotas A, Koniela Z, Wojcik A et al (1994) Coal-bed methane potential of the Upper Silesian Coal Basin. Coal-bed methane potential Up Silesian Coal Basin, Poland, p 142
[29] Kotas A, Porzycki J (1984) Major features of Carboniferous coal basins in Poland. Przegląd Geol 32:268–280
[30] Kuai N, Huang W, Yuan J et al (2012) Experimental investigations of coal dust-inertant mixture explosion behaviors. Procedia Eng 26:1337–1345. https://doi.org/10.1016/j.proeng.2011.11.2309
[31] Küçük A, Kadioǧlu Y, Gülaboǧlu MŞ (2003) A study of spontaneous combustion characteristics of a Turkish lignite: Particle size, moisture of coal, humidity of air. Combust Flame 133:255–261. https://doi.org/10.1016/S0010-2180(02)00553-9
[32] Kusiak MA, Kedzior A, Paszkowski M et al (2006) Provenance implications of Th-U-Pb electron microprobe ages from detrital monazite in the Carboniferous Upper Silesia Coal Basin, Poland. Lithos 88:56–71. https://doi.org/10.1016/j.lithos.2005.08.004
[33] Leonard R, Zulfikar R, Stansbury R (2020) Coal mining and lung disease in the 21st century. Curr Opin Pulm Med 26:135–141. https://doi.org/10.1097/MCP.0000000000000653
[34] Li Q, Lin B, Zhao S, Dai H (2013) Surface physical properties and its effects on the wetting behaviors of respirable coal mine dust. Powder Technol 233:137–145. https://doi.org/10.1016/j.powtec.2012.08.023
[35] Liu G, Vassilev SV, Gao L et al (2005) Mineral and chemical composition and some trace element contents in coals and coal ashes from Huaibei coal field China. Energy Convers Manag. https://doi.org/10.1016/j.enconman.2004.11.002
[36] Liu J, Wang S, Jin L et al (2021) Water-retaining properties of NCZ composite dust suppressant and its wetting ability to hydrophobic coal dust. Int J Coal Sci Technol 8:240–247. https://doi.org/10.1007/s40789-020-00385-2
[37] Markic M, Sachsenhofer RF (1997) Petrographic composition and depositional environments of the Pliocene Velenje lignite seam (Slovenia). Int J Coal Geol 33:229–254. https://doi.org/10.1016/S0166-5162(96)00043-2
[38] Markič M, Sachsenhofer RF (2010) The velenje lignite its petrology and genesis
[39] Marts JA, Gilmore RC, Brune JF, et al (2015) Optimizing nitrogen injection for progessively sealed panels. In: 2015 SME Annual Conference Expo C 117th Natl West Min Conf - Min Navig Glob Waters 447–450
[40] McPherson MJ (1993) Subsurface ventilation systems. Subsurface ventilation and environmental engineering. Springer, Netherlands, pp 91–133
[41] Moreno T, Amato F, Querol X et al (2008) Trace element fractionation processes in resuspended mineral aerosols extracted from Australian continental surface materials. Soil Res 46:128. https://doi.org/10.1071/SR07121
[42] Moreno T, Trechera P, Querol X et al (2019) Trace element fractionation between PM10 and PM2.5 in coal mine dust: Implications for occupational respiratory health. Int J Coal Geol 203:52–59. https://doi.org/10.1016/j.coal.2019.01.006
[43] Palmer AD, Cheng M, Goulet JC, Furimsky E (1990) Relation between particle size and properties of some bituminous coals. Fuel 69:183–188. https://doi.org/10.1016/0016-2361(90)90171-L
[44] dos Pedroso-Fidelis G, S, Farias HR, Mastella GA, et al (2020) Pulmonary oxidative stress in wild bats exposed to coal dust: A model to evaluate the impact of coal mining on health. Ecotoxicol Environ Saf. https://doi.org/10.1016/j.ecoenv.2020.110211
[45] Perret JL, Plush B, Lachapelle P et al (2017) Coal mine dust lung disease in the modern era. Respirology 22:662–670. https://doi.org/10.1111/resp.13034
[46] Querol X, Fernandez Turiel JL, Lopez Soler A, Duran ME (1992) Trace elements in high-S subbituminous coals from the teruel Mining District, northeast Spain. Appl Geochemistry 7:547–561. https://doi.org/10.1016/0883-2927(92)90070-J
[47] Querol X, Whateley MKG, Fernández-Turiel JL, Tuncali E (1997) Geological controls on the mineralogy and geochemistry of the Beypazari lignite, central Anatolia, Turkey. Int J Coal Geol 33:255–271. https://doi.org/10.1016/S0166-5162(96)00044-4
[48] Reed WR, Shahan MR, Zheng Y, Mazzella A (2021) Laboratory results of foam application testing for longwall shield dust control in a simulated environment. Int J Coal Sci Technol 8:217–227. https://doi.org/10.1007/s40789-021-00414-8
[49] Reid DC, Hainsworth DW, Ralston JC, McPhee RJ (2006) Shearer guidance: a major advance in longwall mining. Springer Tracts Adv Robot 24:469–476. https://doi.org/10.1007/10991459_45
[50] Riley KW, French DH, Farrell OP, Wood RA, Huggins FE (2012) Modes of occurrence of trace and minor elements in some Australian coals. Int J Coal Geol 94:214–224
[51] Smith A 6, Diamond WP, Mucho TP, Organiscak A (1994) Bleederless ventilation systems as a spontaneous combustion control measure in U.S. coal mines United States department of the interlor
[52] Song Z, Konietzky H, Herbst M (2020) Drawing mechanism of fractured top coal in longwall top coal caving. Int J Rock Mech Min Sci 130:104329. https://doi.org/10.1016/j.ijrmms.2020.104329
[53] Sperazza M, Moore JN, Hendrix MS (2004) High-resolution particle size analysis of naturally occurring very fine-grained sediment through laser diffractometry. J Sediment Res 74:736–743. https://doi.org/10.1306/031104740736
[54] Stępniewska Z, Pytlak A, Kuźniar A (2014) Distribution of the methanotrophic bacteria in the Western part of the Upper Silesian Coal Basin (Borynia-Zofiówka and Budryk coal mines). Int J Coal Geol 130:70–78. https://doi.org/10.1016/j.coal.2014.05.003
[55] Sunkpal M, Roghanchi P, Kocsis KC (2017) A method to protect mine workers in hot and humid environments. Saf Health Work 9:149–158. https://doi.org/10.1016/j.shaw.2017.06.011
[56] Swaine DJ (1990) Trace elements in coal. Butterworth
[57] Szkudlarek Z, Janas S (2021) Active protection of work area against explosion of dust–gas mixture. Int J Coal Sci Technol. https://doi.org/10.1007/s40789-020-00387-0
[58] Trechera P, Moreno T, Córdoba P et al (2020) Mineralogy, geochemistry and toxicity of size-segregated respirable deposited dust in underground coal mines. J Hazard Mater 399:122935. https://doi.org/10.1016/j.jhazmat.2020.122935
[59] Trechera P, Moreno T, Córdoba P et al (2021) Comprehensive evaluation of potential coal mine dust emissions in an open-pit coal mine in Northwest China. Int J Coal Geol 235:103677. https://doi.org/10.1016/j.coal.2021.103677
[60] Vaněk M, Bora P, Maruszewska EW, Kašparková A (2017) Benchmarking of mining companies extracting hard coal in the Upper Silesian Coal Basin. Resour Policy 53:378–383. https://doi.org/10.1016/j.resourpol.2017.07.010
[61] Więckol-Ryk A, Krzemień A, Sánchez Lasheras F (2018) Assessing the breathing resistance of filtering-facepiece respirators in Polish coal mines: a survey and laboratory study. Int J Ind Ergon 68:101–109. https://doi.org/10.1016/j.ergon.2018.07.001
[62] Woskoboenko F (1988) Explosibility of Victorian brown coal dust. Fuel 67:1062–1068. https://doi.org/10.1016/0016-2361(88)90371-7
[63] Wu Q, Han L, Xu M et al (2019) Effects of occupational exposure to dust on chest radiograph, pulmonary function, blood pressure and electrocardiogram among coal miners in an eastern province. China BMC Public Health 19:1229. https://doi.org/10.1186/s12889-019-7568-5
[64] Yuan J, Wei W, Huang W et al (2014) Experimental investigations on the roles of moisture in coal dust explosion. J Taiwan Inst Chem Eng 45:2325–2333. https://doi.org/10.1016/j.jtice.2014.05.022
[65] Zhang Q, Chen X, Wang H, Xu C (2021) Exploration on molecular dynamics simulation methods of microscopic wetting process for coal dust. Int J Coal Sci Technol 8:205–216. https://doi.org/10.1007/s40789-021-00415-7
[66] Zhang T, Jing D, Ge S et al (2020) Numerical simulation of the dimensional transformation of atomization in a supersonic aerodynamic atomization dust-removing nozzle based on transonic speed compressible flow. Int J Coal Sci Technol 7:597–610. https://doi.org/10.1007/s40789-020-00314-3

Funding

European Commission Research Fund for Coal and Steel (Grant Agreement Number – 754205). Generalitat de Catalunya (SGR41). Centre of Excellence Severo Ochoa—Spanish Ministry of Science and Innovation (Project CEX2018-000794-S).

About this article

Cite this article

Trechera, P., Querol, X., Lah, R. et al. Chemistry and particle size distribution of respirable coal dust in underground mines in Central Eastern Europe.Int J Coal Sci Technol 9, 3 (2022).
  • Received

    24 March 2021

  • Accepted

    04 January 2022

  • DOI

    https://doi.org/10.1007/s40789-022-00468-2

  • Share this article

    Copy to clipboard

For Authors

Explore