Special Issue on Air Pollution and its Impact in South and Southeast Asia (IX)

Zerin Binte Alam This email address is being protected from spambots. You need JavaScript enabled to view it.1,2, Kazi A.B.M. Mohiuddin1 

1 Department of Civil Engineering, Khulna University of Engineering and Technology, Khulna 9203, Bangladesh
2 Department of Civil Engineering, North Western University, Khulna 9100, Bangladesh


Received: May 15, 2022
Revised: September 26, 2022
Accepted: October 11, 2022

 Copyright The Author(s). This is an open access article distributed under the terms of the Creative Commons Attribution License (CC BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are cited.


Download Citation: ||https://doi.org/10.4209/aaqr.220109  

  • Download: PDF


Cite this article:

Alam, Z.B., Mohiuddin K.A.B.M. (2023). Micro-characterization of Dust and Materials of Dust Origin at a Cement Industry Located in Bangladesh. Aerosol Air Qual. Res. 23, 220109. https://doi.org/10.4209/aaqr.220109


HIGHLIGHTS

  • Micro-characterization of cement industrial dust and materials in Bangladesh.
  • Observation of fine and ultra-fine particles with diverse morphology in SEM images.
  • Detection of metallic and non-metallic elements by EDS technique.
  • Finding of minerals with major Portland crystalline phases by XRD, FTIR, SEM-EDS.
  • Revelation of carcinogenic substances such as crystalline silica and asbestoses.
 

ABSTRACT


Industrial dust generation from material processing is an uninterrupted phenomenon, hence analyzing materials of dust origin with dust particles is important to comprehend the entire scenario of occupational health exposure (OHE). In this study, we investigated the morphological, elemental, and mineralogical characteristics of engendered dust and materials of dust origin in the cement industry (CI) in Bangladesh. The dust samples were accumulated from dust collectors and the internal roadways of the CI to understand the nature of atmospheric dust of the CI. All the materials that could potentially be the sources of the dust of the CI were collected, including clinker, gypsum, limestone, fly ash, slag, and two types of cement products. Scanning electron microscopy with energy dispersive x-ray spectroscopy (SEM-EDS), x-ray powder diffraction (XRD), and fourier-transform infrared spectroscopy (FTIR) were performed for the characterization. SEM micrographs showed the presence of fine (particles ≤ PM2.5) and ultra-fine (particles ≤ PM0.1) particles with diverse morphology in the studied samples. Ca, Si, Al, Mg, Fe, K, Na, Mo, and Ti were found as existing metallic elements in samples through the EDS technique. Several minerals of silicate, oxide, carbonate, and sulfate were detected with major crystalline phases of Portland cement by SEM-EDS, XRD, and FTIR analyses in samples; among which some are well documented as occupational hazards. The trace presence of organic carbon was observed in all FTIR spectra. The most significant outcome of this study is the detection of carcinogenic substances such as crystalline silica (quartz, cristobalite) and asbestoses (anthophyllite, chrysotile, and crocidolite) in samples. For instance, quartz was found in the dust of dust collectors while road dust of the CI showed the existence of both quartz and cristobalite. However, asbestoses were found only in source materials, but they could be released anytime during material handling, thus creating OHE.


Keywords: Occupational health exposure, SEM-EDS, XRD, FTIR, Crystalline silica, Asbestos


1 INTRODUCTION


Cement is vital worldwide as a construction material. During cement production, raw materials (limestone, shale, clay, sand, etc.) are ground and heated into a kiln to form the clinker, clinkers are then blended with gypsum or other supplementary cementitious materials, and finally reduced to a fine powder to formulate the cement. This entire process generates enormous pollutants in the atmosphere, creating health and environmental hazards. Typical atmospheric emissions of a cement industry (CI) include oxides of nitrogen, sulfur dioxide, oxides of carbon, dust particles, and some trace elements (Abu-Allaban and Abu-Qudais, 2011). Clinker formation acts as the principal source of gaseous pollutants while dust particles can be generated throughout the cement production process. Quarrying, crushing, and storage of raw materials; clinker production; finish grinding; packaging; and transport of products are reported as possible emission sources of dust generation at CI (Yang et al., 1996; Zhang et al., 2015).

Exposure to cement dust may result in adverse health effects. Several studies stated skin irritation, wheezing, increased mucus production, asthma, bronchiectasis, pneumoconiosis, chronic obstructive pulmonary disease (COPD), preterm delivery, psychasthenia, endocrine disruption, infertility, and carcinoma of the lungs, stomach, and colon as cement dust-related health issues (Yang et al., 1996; Meo, 2004; Tolinggi et al., 2014; Adeyanju and Okeke, 2019).

The health risks at a dusty occupation are primarily linked with the harmful physical, elemental, and mineralogical features of the dust particles (Bickis, 1998). For instance, particles having a diameter ≤ 10 µm can deposit in the upper respiratory tract whereas particles with diameter ≤ 2.5 µm can penetrate the lungs and particles with diameter ≤ 0.1 µm can further reach into the alveolar region of the lungs, respectively (Park and Wexler, 2008; Valiulis et al., 2008). Particle shape is also an influencing factor for particle deposition in respiratory system. Thus, determination of particle morphology is vital.

Dust may contain metallic elements that are potentially hazardous to health (Kicińska, 2016). However, metals in particles can mediate toxicity with their existing chemical features (Oberdörster, 1993; Dick et al., 2003). For example, iron (Fe) is an essential nutrient for human health, but the breathing of Fe-bearing particles can stimulate reactive oxygen species on the lung surface, which may further lead to scarring of the lung tissue (Knaapen et al., 2004; Mohiuddin et al., 2014). Mineralogical structure often distinguishes the degree of toxicity in dust particles, such as, crystalline silica is categorized as group 1 carcinogen by International Agency for Research on Cancer (IARC) and is significantly harmful than amorphous silica on long term inhalation (Sandberg et al., 2012; Suryadevara, 2016). Hence, understanding the elemental distribution and mineralogical structure of dust particles is crucial.

Dust generated at CI may contain hazardous minerals like crystalline silica, asbestiform, and gypsum with metallic elements (Ca, Al, Fe, Mg, etc.) (Neupane et al., 2020). However, occupational exposure to respirable crystalline silica includes silicosis, tuberculosis, chronic bronchitis, and lung cancer (Merget et al., 2002; Calvert et al., 2003). Chronic asbestos exposure may cause two principal types of lung cancer: cancer of the lung tissue itself and mesothelioma, which is a rare form of cancer found in the thin lining of the lung, chest, abdomen, and heart (Koskinen et al., 2002; Berman and Crump, 2008). Asbestos may also cause a serious progressive, long-term, and non-cancer lung disease called asbestosis (Wyers, 1949). Phlegm and dyspnea, irritation of mucous membranes and upper respiratory tract, cough, sneeze, and runny nose have been reported as a result of long-term exposure to gypsum dust (Neghab, 2015). Dust particles bearing metallic elements like Ca, Al, Fe, Mg, etc. may also cause health impacts. Hence, physical, chemical, and mineralogical characterization of dust particles at CI is essential in regard to evaluate their health impacts on workers.

We selected a CI located in Bangladesh to collect the required samples. Bangladesh has a scarcity of mineral resources and hence is highly dependent on imported raw materials including clinker for cement production. Only two companies among forty-two in Bangladesh have clinker production facilities at their plants, the rest of them import materials and use grinding technology to produce cement. The selected CI imports the clinker and the other raw materials as well and produces ordinary portland cement (OPC) and portland composite cement (PCC) as two main cement products. OPC is manufactured here by adding 5% gypsum to 95% clinker whereas 3.5% gypsum, 4% limestone, 8.5% slag, and 22% fly ash are added to 62% clinker for PCC.

Bag filters are installed at required points of CI to capture the generated dust. Beyond the control efficiency of the bag filter or due to poor occupational management during material handling, dust may still be released into the atmosphere and even some of them may deposit on the internal roadways of the CI. Therefore, the study collected dust from both the bag filters and the roadways to understand the nature of atmospheric dust in the CI. Since the dust generation from material processing is a continuous activity, so to comprehend the total scenario of the OHE in the CI we also analyzed the materials (clinker, gypsum, limestone, slag, fly ash, OPC, and PCC) that could be associated with the dust origin. Consequently, the collection and analysis of studies samples were carried out focusing on:

  1. Observation of existing fine and ultra-fine particles with shape diversity in studied samples
  2. Detection of metallic elements in studied samples
  3. Identification of existing crystalline phases, organic and inorganic compounds in studied samples

The outcome of the analysis was the basis of determining the potentially harmful metallic elements, minerals, organic and inorganic compounds in the generated dust and materials of dust origin in the CI of Bangladesh that may cause occupational health hazards unless any safety initiatives were taken. The samples were analyzed using SEM-EDS, XRD, and FTIR techniques.

 
2 METHODS


 
2.1 Study Location

Khulna is the south-western divisional city of Bangladesh. The selected CI is located in the Labanchara area nearby the renowned Rupsha River of Khulna city (Fig. 1). Labanchara is mainly an industrial zone assembled with different industries such as fish processing industries, Bangladesh Oxygen Company, Khulna Shipyard, Bangladesh Match Factory, and so on.

Fig. 1. Labanchara CI in Khulna City (taken from Google Earth).Fig. 1. Labanchara CI in Khulna City (taken from Google Earth).

 
2.2 Sample Collection

The study collected bag filtered dust (BFD) and deposited road dust (DRD) to understand the nature of atmospheric dust at the Labanchara CI. The samples of BFD were accumulated from different bag filters located at the CI, and later were mixed. The samples of DRD were collected from three different of spots of the CI also and then got mixed. Materials that could be associated with the origin of generated dust at the Labanchara CI were gathered for the analysis which include clinker, gypsum, limestone, slag, fly ash, OPC, and PCC. So, total nine samples were collected cautiously in some air tight containers from the Labanchara CI and analyzed by SEM-EDS, XRD, and FTIR techniques. SEM-EDS was performed to obtain the microscopic view of particle morphology with the elemental distribution. XRD analysis was accomplished to identify mineral constituents while FTIR analysis was done to determine compounds both in amorphous and crystalline phases here.

 
2.3 Scanning Electron Microscopic with Energy Dispersive X-ray Spectroscopic Analysis

Scanning electron microscopic with energy dispersive X-ray spectroscopic (SEM-EDS) analysis was carried out using scanning electron microscope (JEOL JSM 6490 LA), equipped with an energy dispersive X-ray spectrometer (JED 2300 Analysis Station). To achieve the SEM micrographs, samples were mounted on aluminum stubs with double-sided sticky carbon tape. A very thin platinum coat was applied to the sample surface to increase the conductivity. The samples were then introduced into the vacuum chamber and subsequently, digital images were obtained. Three images of each sample were taken at different magnifications by stereo SEM at 30 tilts. Elemental analysis was accomplished of bulk samples by detecting characteristic X-rays generated from samples. The dry silicon drift detector of EDS had an acquisition rate of 2504 cps with a resolution of 20 keV. The detection limit of the detector was around 0.1% wt. The ZAF correction was applied to EDS data for the semi-quantification of elements. Here, Z, A, and F refers to the atomic power, absorption correction, and fluorescence correction, respectively.

 
2.4 X-ray Diffraction Analysis

In X-ray diffraction (XRD) analysis, the samples were placed into the ground glass depression of a sample holder and positioned onto the X-ray diffractometer (Rigaku Ultima-IV). The X-ray source was a Cu Kα line with a wavelength of 1.54 Å, and the data were collected at Bragg angle 2θ ranging from 10° to 80°. A specific mineral usually generates a characteristic pattern through XRD analysis. RRUFF™ project provides a complete set of high-quality spectral data including XRD patterns of well-characterized minerals, which can be used as a standard for mineral identification (Lafuente et al., 2015). This study primarily followed the RRUFF database as a standard reference to detect the minerals in samples, by comparing the peak positions at 2θ with corresponding d spacing. In addition, literature studies were followed to confirm a few minerals.

 
2.5 Fourier-Transform Infrared Spectroscopic Analysis

For fourier-transform infrared spectroscopic (FTIR) analysis, collected ground samples were mixed with an infrared transparent soft salt i.e., potassium bromide (KBr), with a ratio of 20:1. The mixer was then pelletized using hydraulic pressure of 10 tons. KBr based pellets were then analyzed using Shimadzu FTIR-8400s spectrophotometer. The spectral resolution of the spectrometer was 4 cm–1. The infrared (IR) spectra were obtained in the range of 4000–500 cm–1, with an average of 32 scans. The data were collected in attenuated total reflectance (ATR) mode. Identification of minerals and compounds was done accordant with literature studies.

 
3 RESULTS AND DISCUSSION


 
3.1 SEM-EDS Analysis

Particles with diverse morphology varying from 0.1–15 µm were observed in SEM micrographs of studied samples (Figs. 2–5) here. EDS analysis simultaneously detected Ca, Si, Al, Mg, Fe, Mo, Ti, K, Na as metallic elements and O, C, S as non-metallic elements, as shown in Table 1.

Fig. 2. (a–b) SEM micrographs of gypsum and limestone.Fig. 2. (a–b) SEM micrographs of gypsum and limestone.

Table 1. Semi-quantitative Elemental Analysis of Samples by SEM-EDS technique.

Particles with tabular, prismatic, acicular, or massive nature viewed in the micrograph of gypsum (Fig. 2(a)) might indicate the gypsum minerals (CaSO4.2H2O) while silica (SiO2) is observed as clustered form dispersed in between gypsum (Ausset et al., 2000; Shih et al., 2005; Ashrit et al., 2015). Irregular-angular to quasi-spherical shape with rough surface seen in Fig. 2(a) can be the particles of alumina (Al2O3) and the cubic or rectangular shape might be signifying the molybdenum metal (Mo) (Laha et al., 2005; Yehia et al., 2020). Calcite (CaCO3) and dolomite (CaMg(CO3)2) are the primary minerals of limestone (Zhang and Lv, 2020). Particles of both calcite and dolomite possess layered rhombohedral shape under the microscope, which are abundantly seen in the SEM image of limestone (Fig. 2(b)) (Yip et al., 2008; Kasha et al., 2015). Small round and angular-shaped particles might be indicative of silica and alumina, respectively, as they may exist as impurities in limestone (Dey et al., 2020). Fine and ultrafine particles are spotted in Figs. 2(a) and 2(b). The elemental distribution of gypsum and limestone is consistent with these findings (Table 1).

The predominant presence of microspheres is found in the SEM images of fly ash (Figs. 3(a–b)). Microspheres in fly ash result from the sintering of ash at high combustion temperature, and alumino‐silicate glass, mullite, quartz, iron oxides, calcium sulfates, etc. are usually observed as their major phases (Żyrkowski et al., 2016; Zong et al., 2019). These microspheres may also consist of Fe alloys with admixtures of Pb and Ba; or are sometimes composed of entirely of carbon (Kicińska and Bożęcki, 2018). The irregular grains with hollow features primarily indicate mullite-sillimanite minerals (3Al2O3.2SiO2) (Tazaki et al., 1989; Zong et al., 2019). The particle of nearly hexagonal shape viewed in fly ash micrograph (Fig. 3(a)) might be signifying kaolinite (Al2Si2O5(OH)4) (Tazaki et al., 1989). K and Ti found in elemental analysis of sampled fly ash, were commonly reported as minor elements of coal fly ash in earlier studies (Martinez-Tarazona and Spears, 1996; Rautray et al., 2009; Tiwari et al., 2014).

Fig. 3. (a–d) SEM micrographs of fly ash and slag.Fig. 3. (a–d) SEM micrographs of fly ash and slag.

Ground granulated blast-furnace slag (GGBFS) is used in the CI due to its high cementitious properties. The presence of Ca, Si, Al, Mg, O as major elements in EDS data implies the sampled slag is a blast furnace slag (Grubeša et al., 2016). Glassy or crystalline phases of gehlenite Ca₂Al(AlSi)O₇), akermanite (Ca₂Mg(Si₂O₇)), merwinite (Ca3Mg(SiO4)2), rankinite (Ca3Si2O7), and wollastonite (CaSiO3) are to be generally present in GGBFS (Snellings et al., 2012; Tripathy et al., 2020). Akermanite, rankinite, and wollastonite can be spotted by the structure with elongated, plate-like, and stick shapes, respectively, while gehlenite with irregular shapes, in slag micrograph (Fig. 3(a)) (Sofilić et al., 2013; Wilczyńska-Michalik et al., 2015). Figs. 3(b) and 3(d) are showing particles with diameter ranging from 10 to 0.1 µm. The elemental distribution of fly ash and slag has shown similar findings (Table 1). However, molybdenum found in the sample of slag is considered as the most valuable element that can exist in slag (Parada et al., 2009).

Alite (Ca3SiO5), belite (Ca2SiO4), aluminate (Ca3Al2O6), and ferrite (Ca4Al2Fe2O10) are the typical mineral phases of portland cement. The small roundish particles seen in microscopic images of clinker, PCC, and OPC (Figs. 4(a–c)) are the belite minerals while euhedral to anhedral-shaped particles usually bigger than belite are the alite minerals (Kolovos et al., 2005; Stutzman et al., 2016). The elongated or prismatic particles indicate ferrite phases whereas aluminate phases are seen as fine to lath-like crystals consisting of a matrix between ferrite crystals (Stutzman et al., 2016). Periclase and alkali sulfates (arcanite, thenardite, etc.) might appear in the dark zone like the voids (Stutzman et al., 2016). Free lime might be seen as bright small round shapes or sometimes as clustered form (Stutzman et al., 2016). Asbestos is the fibrous silicate mineral and particles with acicular shape seen in OPC might be indicating asbestos. SEM images of PCC basically presents particles with different morphology originated from the supplementary materials (slag, fly ash, limestone) in addition to clinker and gypsum.

Fig. 4. (a–c) SEM micrographs of clinker, OPC, and PCC.Fig. 4. (a–c) SEM micrographs of clinker, OPC, and PCC.

Apparently, the BFD micrograph (Fig. 5(a)) exhibited irregularly fume-shaped particles. The particles of BFD might be very fine, to be detected with their actual morphology in microscales. EDS analysis found Ca, Si, Al, Fe, Mg, Mo, K, O, and C as existed elements in BFD. Particles with irregular and rounded shapes observed in the DRD micrograph (Fig. 5(b)) might be indicating soil-derived minerals and quartz. Kaolinite might be seen as the nearly hexagonal shape ranges more than 15 microns in Fig. 5(b). EDS data showed the abundance of Si, Al, and O elements in DRD. The presence of a significant amount of Na may indicate clayey minerals such as albite, beidellite.

Fig. 5. (a–b) SEM micrographs of BFD and DRD.Fig. 5. (a–b) SEM micrographs of BFD and DRD.


3.2 XRD Analysis

XRD is one of the highly applied techniques to determine the mineralogical composition in materials. The minerals detected in studied samples by XRD analysis are presented in Table 2 with their corresponding diffraction peaks (2θ values) and d-spacings. Fig. 6 exhibits the XRD diffractograms of all collected samples.

Fig. 6. XRD diffractograms of studied samples.Fig. 6. XRD diffractograms of studied samples.

Table 2. Mineralogical Analysis of Studied Samples by XRD.

The silicate minerals are the primary concern because of their relative abundance and significance (Kumar and Rajkumar, 2014). Quartz, mullite-sillimanite, rankinite, wollastonite, albite, beidellite, kaolinite, microcline, anthophyllite, and serpentine minerals were found as existing silicate minerals here along with Portland cement-based calcium silicate phases (alite, belite, aluminate, and ferrite). Alite, belite, aluminate, and ferrite were detected in OPC, PCC, clinker, and BFD (Jadhav and Debnath, 2011; Sedaghat et al., 2014). Anthophyllite is an amphibole mineral from the inosilicate group, contains the chemical formula: ☐Mg2Mg5Si8O22(OH)2 (☐ is for a vacancy, a point defect in the crystal structure), might also be present in four of them (Lafuente et al., 2015). Serpentine is the group of hydrous magnesium-rich phyllosilicate minerals, and a diffraction peak at 12.18° at 2θ in the XRD pattern of OPC indicates the existence of serpentine minerals (Lafuente et al., 2015). Chrysotile, antigorite, and lizardite are three of the primary serpentine minerals and their composition approximates Mg3Si2O5(OH)4.

The most common form of crystalline silica, quartz was detected in samples of PCC, gypsum, fly ash, limestone, BFD, and DRD (Lafuente et al., 2015; Neupane et al., 2020). Cristobalite is another polymorph of crystalline silica that is formed at very high temperatures, found to be present in DRD (Yusan et al., 2012; Lafuente et al., 2015). Experimental evidence shows cristobalite and quartz as more cytotoxic than other polymorphic forms of crystalline silica (Meldrum and Howden, 2002). The XRD diffractogram of fly ash revealed the abundance of mullite-sillimanite minerals (Jozić et al., 2010; Lafuente et al., 2015). Both sillimanite and mullite have similar d-spacings but their compositions are different; sillimanite is Al2O3·SiO2 while mullite is 1.6–1.9 Al2O3·SiO2 (Tazaki et al., 1989). The respiratory hazard of mullite or sillimanite is little known, although one study cited chronic bronchitis, silico-tuberculosis, and pneumoconiosis as occupational exposure of mullite (Artamonova et al., 2000). In addition, Brown et al. (2011) stated mullite as a potential respiratory hazard in their study.

Typical GGBFS minerals i.e., akermanite, gehlenite, rankinite, merwinite, and wollastonite usually produce major diffraction peaks between 20°–40° at 2θ range during XRD analysis (Kasina et al., 2014; Lafuente et al., 2015). The presence of a broad hump in the range of 20°–40° at 2θ with only two discrete maxima indicates the glassy nature of sampled slag. The distinct peak at 30.5° at 2θ in slag XRD pattern might be attributed to akermanite or rankinite, or wollastonite crystals. Albite (NaAlSi₃O₈) and microcline (KAlSi₃O₈) are Na-rich plagioclase feldspar and K-rich alkali feldspar, respectively, and belong to the tectosilicate mineral group. The obvious presence of albite was found in DRD with a trace amount of microcline (Lafuente et al., 2015). Two clayey minerals from the phyllosilicate group i.e., beidellite and kaolinite were also detected in DRD (Lafuente et al., 2015; Neupane et al., 2020). Beidellite contains chemical formula as (Na,Ca)0.3Al2(Si,Al)4O10(OH)2·nH2O.

The study found the existence of dolomite and three main polymorphs of calcium carbonates i.e., vaterite, aragonite, calcite, as carbonate minerals in different samples. Calcite is the most stable polymorph of calcium carbonate found as governing mineral in the sampled limestone (Jadhav and Debnath, 2011; Lafuente et al., 2015). The samples of OPC, PCC, clinker, and BFD also indicated the existence of calcite. Inhalation of respirable calcite can irritate eyes, skins, lungs, and its higher exposure may cause pulmonary edema (Manisalidis et al., 2020).  Aragonite is preferred phase of calcium carbonates at high pressures and low temperatures and was detected only in BFD (Lafuente et al., 2015). The high pressure of the bag-filter might cause the alteration of calcite into aragonite. The presence of vaterite and dolomite was observed in DRD and limestone, respectively (Lafuente et al., 2015; Stutzman et al., 2016). A case study by Mishra et al. (2004) stated that prolonged exposure to dust in mines of limestone and dolomite might have caused tuberculosis in several workers.

The occurrence of gypsum mineral was evidently identified in the sample of gypsum (Jadhav and Debnath, 2011; Stutzman et al., 2016). Mineral of gypsum was also found in OPC, PCC, clinker, and BFD. Arcanite (K₂SO₄) is a sulfate mineral, usually known as alkali when exists in cementitious materials. The presence of arcanite was noticed in PCC, fly ash, and BFD (Lafuente et al., 2015; Stutzman et al., 2016).

Corundum (Al2O3), periclase (MgO), hematite (Fe2O3), brookite (TiO2), and molybdite (MoO3) were the found as oxide minerals in samples. The crystalline form of aluminum oxide, known as corundum, was observed in gypsum, limestone, BFD, and DRD (Lafuente et al., 2015). The samples of OPC, PCC, clinker, and BFD showed the presence of both periclase and hematite (Behera and Sarkar, 2016; Cvetković et al., 2018). Periclase and hematite were also detected in slag and fly ash, respectively. Wüstite (FeO) is another mineral form of iron oxide like hematite and might exist in slag (Shao et al., 2015). Breathing of alumina dust can generate local inflammation in the respiratory tract which may lead to occupational asthma and fibrosis (Haleatek et al., 2005). Inhaling magnesia and iron oxides fume can cause metal fume fever flu-like illness with symptoms of cough, fever, chills, malaise, and myalgias (Drinker et al., 1927; Mueller and Seger, 1985). Chronic exposure to iron oxides fume might instigate fibrotic pulmonary changes (Jones and Warner, 1972).

Brookite (TiO2) is the orthorhombic variant of titanium dioxide, and the characteristic peaks of brookite were observed in the XRD pattern of fly ash (Lafuente et al., 2015). XRD diffractogram of BFD showed the presence of molybdite (MoO3), a natural mineral form of molybdenum trioxide (Alemán-Vázquez et al., 2005; Lafuente et al., 2015). In addition, the study found the occurrence of molybdenum metal in gypsum (Lafuente et al., 2015; Pandharkar et al., 2018). Molybdenum trioxide is classified as possible carcinogen (Group 2B) while high concentrations of molybdenum metal may irritate the upper respiratory tract and cause a gout-like syndrome in the human body (Koval'skii et al., 1961; Barceloux and Barceloux, 1999). In addition, several studies reported that nanoparticles of titania accumulate in the lungs, alimentary tract, liver, heart, spleen, kidneys, and cardiac muscle after inhalation or oral exposure (Wójcik et al., 2020).

 
3.3 FTIR Analysis

FTIR is a universal analytical tool used to evaluate a wide range of materials, especially to detect unknown materials. The study found the existence of several minerals, compounds with amorphous or glassy nature, and organic carbon through FTIR analysis. The vibrational frequencies in the IR spectra of studied samples and their tentative assignment to the corresponding minerals and compounds are presented in Table 3. Fig. 7 are presenting IR spectra of studied samples.

Table 3. Vibrational frequencies of studied samples and their tentative assignments.

Table 3. (continued).

Fig. 7. IR spectra of studied samples.Fig. 7. IR spectra of studied samples.

IR spectra of OPC and clinker, showing medium intense broad peaks at 927 and 931 cm1, respectively, might be occurred from asymmetric stretching vibrations of Si-O-Si in SiO4 tetrahedra of alite (Ghosh and Handoo, 1980; Fernández-Carrasco et al., 2012). Another medium intense peak with sharp nature at 522 cm1, viewed in both OPC and clinker, can be produced from out-of-plane bending vibrations of Si-O in SiO4 tetrahedra of alite or belite (Ghosh and Handoo, 1980; Fernández-Carrasco et al., 2012). The other two vibrational peaks seen at 457 and 460 cm1 in OPC and clinker, respectively, might be resulted from the in-plane Si-O bending in SiO4 tetrahedra of alite (Ghosh and Handoo, 1980; Fernández-Carrasco et al., 2012). These two vibrational peaks (522 and 460 cm1) can also be resulted from the bending of Al-O bonds within AlO6 octahedral groups present in aluminate (Ghosh and Handoo, 1980; Fernández-Carrasco et al., 2012).

The pure phases of alite and belite usually generate vibrations in the range of 1000-800 cm1 due to the stretching of Si-O-Si in their SiO4 tetrahedra (Ghosh and Handoo, 1980; Fernandez-Carrasco et al., 2012). Although, the presence of impurities (MgO, Na2O, Al2O3, Fe2O3) triggers a change in the crystalline structure of the silicate phases which may cause modifications in the IR spectra (Fernandez-Carrasco et al., 2012). IR spectrum of PCC showing the main band near 1080 cm1 with strong and broad nature, might be originated from the vibrations of calcium silicate phases i.e., alite, belite. The vibrational peak at 1080 cm1 may also indicate the presence of quartz, sulphates, and amorphous silica in PCC (Ojima, 2003; Fernandez-Carrasco et al., 2012). Two sharp vibrational peaks at 779 and 462 cm1 in the PCC spectrum can result from stretching of AlO4 tetrahedra and bending of AlO6 octahedra in aluminates, respectively, while another one at 688 cm1 might be derived from OH liberation modes of anthophyllite (Fernandez-Carrasco et al., 2012; Della Ventura et al., 2018). These three peaks (779, 688, 462 cm1) can be occurred from stretching and bending vibrations of Si-O in SiO4 tetrahedra of quartz as well (Ojima, 2003; Kumar and Rajkumar, 2014). However, the sharp and single peak at 779 cm1, has also been reported as characteristic peak for crocidolite (Della Ventura et al., 2018; Neupane et al., 2020).

IR spectrum of BFD exhibiting vibrational peak at 997 cm1, might be attributing the asymmetric stretching of Si-O-Si of belite (Ghosh and Handoo, 1980; Fernandez-Carrasco et al., 2012). Orthorhombic crystal unit of molybdenum trioxide may also produce band at 997 cm1 due to unresolved stretching vibrations of Mo=O (Seguin et al., 1995; Ding et al., 2006). Another peak at 869 cm1 in the BFD spectrum, can occur from the Al-O stretching in AlO4 tetrahedra of aluminate, or from symmetric Si-O-Si stretching in SiO4 tetrahedra of alite (Ghosh and Handoo, 1980; Fernandez-Carrasco et al., 2012). However, stretching of the oxygen atoms in a Mo-O-Mo entity of molybdenum trioxide may generate vibrations near 869 cm1 (Seguin et al., 1995). A broad natured peak seen at 470 cm1 in the BFD spectrum is characteristic for hydrated amorphous silica (Fröhlich, 1989; Bertaux et al., 1998). A shoulder peak at 1107 cm1 in clinker, and more distinctive peaks at 1111 cm1 in OPC and 1109 cm1 in BFD, might be indicating the presence of sulphates, hematite, and anthophyllite (Li et al., 2011; Fernández-Carrasco et al., 2012). However, FTIR analysis could not find any significant frequencies for ferrite in samples of OPC, PCC, clinker, and BFD which is consistent with the earlier study (Hughes et al., 1995).

A broad and strong peak at 3441 cm1 in the OPC spectrum might be originated from stretching vibrations of the Si-OH bond (Soheilmoghaddam et al., 2014; Jiang et al., 2017). PCC, clinker, BFD showing spectra at 3435, 3450, and 3454 cm1, respectively, can be derived from OH stretching vibrations in molybdenum trioxide (Seguin et al., 1995). Chrysotile consists of octahedral sheets of magnesium hydroxide (brucite) covalently bonded to tetrahedral sheets of silicon oxide (tridymite), with a regular layered structure (Falini et al., 2004). The characteristic peak of chrysotile resulting from external Mg-OH stretching in octahedral sheets of brucite, can be observed at 3699 cm1 in the IR spectrum of OPC (Falini et al., 2004).

A vibrational peak near 800 cm1 occurred from Si-O-Si symmetrical stretching is mostly used as an analytical peak of crystalline silica (Ojima, 2003; Hart et al., 2018). Quartz, cristobalite, and tridymite are the main three polymorphs of crystalline silica. Vibrational peaks at 464 and 1089 cm1 in the IR spectrum of fly ash might have resulted from bending and asymmetric stretching of Si-O, respectively, indicating the presence of quartz, cristobalite, and amorphous silica (Ojima, 2003; Correcher et al., 2009; Neupane et al., 2020). A small and single peak at 790 cm1 in the fly ash spectrum might have occurred from the presence of crystalline silica i.e., cristobalite or tridymite (Correcher et al., 2009). Asymmetric stretching of S-O in arcanite and asymmetric stretching of (Si, Al)-O-Si in mullite may also generate vibrations near 1089 cm1 (Ojima, 2003; Zhang et al., 2012). A sharp medium intense peak at 551 cm1 might have occurred from Al-O stretching of mullite or Si-O bending of cristobalite (Ojima, 2003). However, a vibrational peak near 551 cm1 resulting from Fe-O stretching is also typical for hematite (Li et al., 2019). A very weak peak observed near 850 cm1 in the fly ash spectrum might be generated from the stretching of Ti-O-Ti in the titania polymorph (Kralevich and Koenig, 1998). Another sharp medium intense peak at 1631 cm1 might have resulted from O-H bending vibrations and characteristics for both mullite and titania polymorph (Ojima, 2003; Zhang et al., 2017). A strong peak with broad nature at 3454 cm1 might be generated from OH stretching vibrations, suggesting the presence of silicate material (Escribano et al., 2017). IR spectrum of fly ash shows a weak peak at 3695 cm1 which may imply the presence of hydroxyl group in kaolinite or alumina (Ojima, 2003; Kumar and Rajkumar, 2014). Also, the peak at 3691 cm1 in the PCC spectrum might have originated from kaolinite of fly ash.

Vibrational peaks at 601 and 669 cm1 in the gypsum spectrum can be assigned to S-O bending and 1126 cm1 to asymmetric S-O-S stretching (Fernández-Carrasco et al., 2012). Another peak at 466 cm1 might have occurred from bending modes of Si-O of quartz (Ojima, 2003). IR spectrum of gypsum exhibiting two sharp peaks at 1622 and 1685 cm1 can be produced from O-H bending vibrations (Fernández-Carrasco et al., 2012). The other two strong broad peaks at 3404 and 3545 cm1 might be attributed to the stretching of the OH group. A very weak peak centered at 3770 cm1 might be indicating a free active hydroxyl group that characterizes alumina in gypsum (Escribano et al., 2017). Accordingly, IR spectra of OPC, PCC, clinker, fly ash, slag, limestone, BFD, DRD demonstrated peaks at 3776, 3770, 3774, 3772, 3776, 3772, 3774, 3770 cm1, respectively might be resulted from hydroxyl groups of alumina.

A broad and strong vibrational peak at 997 cm1 in the IR spectrum of slag might have generated from asymmetric stretching of Si(Al)-O of glass gehlenite (Taylor, 1990). The glass akermanite may also produce a vibrational peak near 997 cm1 due to the stretching of Si-O-Si (Dowty, 1987). However, crystalline phases of rankinite and belite generate characteristic vibrational peaks near 997 cm1 due to the asymmetric stretching of Si-O-Si (Fernández-Carrasco et al., 2012; Wang et al., 2018). A shoulder peak at 881 cm1 in the slag spectrum might have occurred from the presence of an appreciable amount of SiO44– groups in the glass network of gehlenite or from the oscillation of oxygen atoms between Si and Mg in akermanite (Sharma et al., 1983; Dowty, 1987). IR spectrum of slag showing a weak peak at 673 cm1 might be attributed to Si-O-Si symmetric stretching of wollastonite or Al-O-Al symmetric stretching of gehlenite (Taylor, 1990; Ding et al., 2014). A medium intense peak at 474 cm1 resulting from bending of Si-O can be originated from crystalline phases of both wollastonite and akermanite (Kimata, 1980; Ding et al., 2014). Epidemiological evidence of wollastonite demonstrated a nonspecific increase in bronchitis and reduced lung function (Maxim and McConnell, 2005). A broad and strong intense peak at 3446 cm1 occurring from OH stretching might be an indicator of amorphous silica (Ojima, 2003).

Two strong vibrational peaks at 925 and 881 cm1 in the DRD spectrum might have resulted from the OH liberation mode associated with octahedral Al-OH groups in beidellite structure (Farmer, 1974; Kloprogge and Frost, 1999). The peak centered at 520 cm1 might be presenting the coupling between the O-Si-O deformation and the Na-O stretching due to albite in DRD (Dowty, 1987; Zhang et al., 2019). The vibrational peak at 1101 cm1 might have derived from asymmetric stretching of Si-O-Si of albite or cristobalite while at 459 cm1 due to Si-O bending of quartz (Mollah et al., 1992; Ojima, 2003). The peaks at 881 and 1431 cm1 may occur from vibrations of carbonate bonds that existed in vaterite (Neupane et al., 2020). A weak vibrational peak at 3691 cm1 in the DRD spectrum may indicate the presence of the O-H group of kaolinite (Ojima, 2003; Kumar and Rajkumar, 2014).

Vibrational peaks at 1427, 875, and 709 cm–1 in the IR spectrum of limestone can be assigned to the asymmetric stretching, out-of-plane bending, and in-plane bending mode of CO32, respectively (Hsiao et al., 2019). The vibrational peaks at 1427 cm–1 and 875 cm–1, are characteristic for both calcite and dolomite, although the small sharp peak at 709 cm-l distinguishes calcite from dolomite (Reig et al., 2002; Hsiao et al., 2019). The weak peak appearing at 1801 cm–1 in the limestone spectrum is also an indication of the presence of CO32 (Nagabhushana et al., 2008; Hsiao et al., 2019). Accordingly, the vibrational peaks with broad nature appearing at 1438 cm–1 in OPC, 1409 cm–1 in PCC, and 1433 cm−1 in clinker might be attributed to the asymmetric stretching of CO32 of calcite. However, a medium intense peak at 1477 cm–1 in the BFD spectrum, might have arisen from asymmetric stretching of CO32− of aragonite (Toffolo et al., 2019).

Two weak peaks at 1373 and 1423 cm−1 in the slag spectrum might be corresponding to the asymmetric stretching vibration of O-C-O bonds indicating a slight degree of carbonation that has already taken place in the raw material (Gao et al., 2014). Another weak vibrational peak at 1379 cm−1 detected in the fly ash spectrum might also be attributed to CO32– probably due to some carbonation during the sample preparation (Rafeet et al., 2019). In addition, the limestone spectrum shows vibrational peaks at 1111 and 474 cm1 which might have resulted from quartz. The very weak peak at 3697 cm1 might be generated from hydroxyl groups of Mg(OH)2 (Wang et al., 2016; Jing et al., 2019). Vibrational frequencies between 2850–2950 are ascribed to the stretching mode of C-H bonds, found in all samples. IR spectra at 2935, 2927, 2926, 2924 cm1 and 2868, 2864, 2862, 2860, 2858, 2854, and 1705 cm1, are attributed to asymmetric and symmetric stretching of C-H, respectively (Kumar and Rajkumar, 2014).

 
4 CONCLUSION


  • SEM micrograph of BFD exhibited roundish and irregularly fume-shaped particles whereas DRD showed particles with irregular, hexagonal, and roundish shapes. Other studied samples i.e., raw materials and final products demonstrated diverse particle morphology, such as tabular, prismatic, rectangular, angular, irregular, spherical, quasi-spherical, acicular, platy, rhombohedral, hexagonal, elongated, and roundish; among which microparticles with acicular or fibrous shapes are perilous for health as they can travel further in the lung airway.

  • Samples of DRD and BFD showed particles ranging from 0.1 µm to nearly 15 µm through SEM analysis. Fine and ultra-fine particles were also seen here in micrographic views of the rest of the samples.

  • EDS analysis detected Ca, Si, Al, Mg, Fe, Mo, Ti, K, Na as metallic elements and O, C, S as non-metallic elements in studied samples, distributed with different percentages.

  • The presence of silicate minerals (quartz, cristobalite, mullite-sillimanite, kaolinite, beidellite, akermanite, gehlenite, rankinite, wollastonite, albite, anthophyllite, chrysotile, crocidolite); carbonate minerals (calcite, aragonite, vaterite, dolomite); sulfates minerals (gypsum, arcanite); oxides minerals (alumina, hematite, wüstite, periclase, brookite, molybdite); and molybdenum were found in studied samples along with Portland cement-based minerals, by SEM-EDS, XRD, and FTIR techniques. Breathing some of these minerals may cause health hazards unless any safety measures are taken.

  • The occurrence of organic carbon and a few silicate compounds such as amorphous silica, glass gehlenite, glass akermanite, opal was viewed in the IR spectra of different samples.

  • The occupational health hazard of carcinogens like crystalline silica (quartz, cristobalite) is well recognized. Quartz was found in BFD, while DRD showed the existence of both quartz and cristobalite. A potential carcinogen, molybdenum trioxide was also detected in the sample of BFD.

  • Another carcinogen: asbestoses (anthophyllite, chrysotile, crocidolite) were found in the samples of OPC, PCC, and clinker, but not in the BFD and DRD. But workers could be exposed anytime to asbestoses during material handling, or even at the construction site, where the cement products would be used.

  • Therefore, this study recommends Government of Bangladesh to implementing strict laws for occupational health safety in CIs regarding the potential toxicity of generated dust.

 
ACKNOWLEDGEMENTS


The authors would like to express their gratitude to the Khulna University of Engineering & Technology to provide financial aid to this research project. They also want to give appreciation to the University of Dhaka to permit to conduct the tests.


REFERENCES


  1. Abu-Allaban, M., Abu-Qudais, H. (2011). Impact assessment of ambient air quality by cement industry: A case study in Jordan. Aerosol Air Qual. Res. 11, 802–810. https://doi.org/10.4209/​aaqr.2011.07.0090

  2. Adeyanju, E., Okeke, C.A. (2019). Exposure effect to cement dust pollution: A mini review. SN Appl. Sci. 1, 1572. https://doi.org/10.1007/s42452-019-1583-0

  3. Alemán-Vázquez, L., Torres-García, E., Villagómez-Ibarra, J., Cano-Domínguez, J. (2005). Effect of the particle size on the activity of MoOXCY catalysts for the isomerization of heptane. Catal. Lett. 100, 219–226. https://doi.org/10.1007/s10562-004-3459-0

  4. Artamonova, V., Fishman, B., Lashina, E., Medik, V., Veber, V. (2000). Clinical features of respiratory diseases caused by mullite dust. Med. Tr. Prom. Ekol. 10, 17–21. PMID: 11109783.

  5. Ashrit, S., Banerjee, P.K., Chatti, R.V., Venugopal, R., Udayabhanu, G. (2015). Characterization of gypsum synthesized from LD slag fines generated at a waste recycling plant of a steel plant. New J. Chem. 39, 4128–4134. https://doi.org/10.1039/C4NJ02023E

  6. Ausset, P., Lefèvre, R.A., Del Monte, M. (2000). Early mechanisms of development of sulphated black crusts on carbonate stone, in: Fassina, V. (Ed.), Proceedings of the 9th International Congress on Deterioration and Conservation of Stone, Elsevier Science B.V., Amsterdam, pp. 329–337. https://doi.org/10.1016/B978-044450517-0/50115-X

  7. Barceloux, D.G., Barceloux, D. (1999). Molybdenum. J. Toxicol. Clin. Toxicol. 37, 231–237. https://doi.org/10.1081/clt-100102422

  8. Behera, S., Sarkar, R. (2016). Effect of different metal powder anti-oxidants on N220 nano carbon containing low carbon MgO-C refractory: An in-depth investigation. Ceram. Int. 42, 18484–18494. https://doi.org/10.1016/j.ceramint.2016.08.185

  9. Berman, D.W., Crump, K.S. (2008). Update of potency factors for asbestos-related lung cancer and mesothelioma. Crit. Rev. Toxicol. 38, 1–47. https://doi.org/10.1080/10408440802276167

  10. Bertaux, J., Froehlich, F., Ildefonse, P. (1998). Multicomponent analysis of FTIR spectra; quantification of amorphous and crystallized mineral phases in synthetic and natural sediments. J. Sediment. Res. 68, 440–447. https://doi.org/10.2110/jsr.68.440

  11. Bickis, U. (1998). Hazard Prevention and Control in the Work Environment: Airborne Dust. World Health 13, 16.

  12. Brown, P., Jones, T., BéruBé, K. (2011). The internal microstructure and fibrous mineralogy of fly ash from coal-burning power stations. Environ. Pollut. 159, 3324–3333. https://doi.org/​10.1016/j.envpol.2011.08.041

  13. Calvert, G., Rice, F., Boiano, J., Sheehy, J., Sanderson, W. (2003). Occupational silica exposure and risk of various diseases: An analysis using death certificates from 27 states of the United States. Occup. Environ. Med. 60, 122–129. http://dx.doi.org/10.1136/oem.60.2.122

  14. Correcher, V., Garcia-Guinea, J., Bustillo, M., Garcia, R. (2009). Study of the thermoluminescence emission of a natural α-cristobalite. Radiat. Eff. Defects Solids 164, 59–67. https://doi.org/​10.1080/10420150802270995

  15. Cvetković, V.S., Vukićević, N.M., Nikolić, N.D., Branković, G., Barudžija, T.S., Jovićević, J.N. (2018). Formation of needle-like and honeycomb-like magnesium oxide/hydroxide structures by electrodeposition from magnesium nitrate melts. Electrochim. Acta 268, 494–502. https://doi.org/10.1016/j.electacta.2018.02.121

  16. Della Ventura, G., Vigliaturo, R., Gieré, R., Pollastri, S., Gualtieri, A.F., Iezzi, G. (2018). Infra red spectroscopy of the regulated asbestos amphiboles. Minerals 8, 413. https://doi.org/10.3390/​min8090413

  17. Dey, S., Sahu, L., Chaurasia, B., Nayak, B. (2020). Prospects of Utilization of waste dumped low-grade limestone for iron making: A case study. Int. J. Min. Sci. Technol. 30, 367–372. https://doi.org/10.1016/j.ijmst.2020.03.011

  18. Dick, C.A., Brown, D.M., Donaldson, K., Stone, V. (2003). The role of free radicals in the toxic and inflammatory effects of four different ultrafine particle types. Inhalation Toxicol. 15, 39–52. https://doi.org/10.1080/08958370304454

  19. Ding, Q., Huang, H., Duan, J., Gong, J., Yang, S., Zhao, X., Du, Y. (2006). Molybdenum trioxide nanostructures prepared by thermal oxidization of molybdenum. J. Cryst. Growth. 294, 304–308. https://doi.org/10.1016/j.jcrysgro.2006.07.004

  20. Ding, W., Fu, L., Ouyang, J., Yang, H. (2014). CO2 mineral sequestration by wollastonite carbonation. Phys. Chem. Miner. 41, 489–496. https://doi.org/10.1007/s00269-014-0659-z

  21. Dowty, E. (1987). Vibrational interactions of tetrahedra in silicate glasses and crystals. Phys. Chem. Miner. 14, 542–552. https://doi.org/10.1007/BF00308290

  22. Drinker, P., Thomson, R.M., Finn, J.L. (1927). Metal fume fever: III. The effects of inhaling magnesium oxide fume. J. Ind. Hyg. 9, 187–192. https://www.cabdirect.org/cabdirect/​abstract/19272701396

  23. Escribano, V.S., Garbarino, G., Finocchio, E. (2017). γ-Alumina and amorphous silica–alumina: Structural features, acid sites and the role of adsorbed water. Top. Catal. 60, 1554–1564. https://doi.org/10.1007/s11244-017-0838-5

  24. Falini, G., Foresti, E., Gazzano, M., Gualtieri, A.F., Leoni, M., Lesci, I.G., Roveri, N. (2004). Tubular-shaped stoichiometric chrysotile nanocrystals. Chem. Eur. J. 10, 3043–3049. https://doi.org/​10.1002/chem.200305685

  25. Farmer, V.C. (Ed.) (1974). The Infrared Spectra of Minerals. Mineralogical Society of Great Britain and Ireland. https://doi.org/10.1180/mono-4

  26. Feng, D., Provis, J.L., Deventer, J.S. (2012). Thermal activation of albite for the synthesis of one‐part mix geopolymers. J. Am. Ceram. Soc. 95, 565–572. https://doi.org/10.1111/j.1551-2916.2011.04925.x

  27. Fernández-Carrasco, L., Torrens-Martín, D., Morales, L.M., Martínez-Ramírez, S. (2012). Infrared Spectroscopy in the Analysis of Building and Construction Materials, in: Theophanides, T. (Ed.), Infrared Spectroscopy - Materials Science, Engineering and Technology, InTech. pp. 369–382. https://doi.org/10.5772/36186

  28. Fröhlich, F. (1989). Deep‐sea biogenic silica: New structural and analytical data from infrared analysis‐geological implications. Terra. Nova 1, 267–273. https://doi.org/10.1111/j.1365-3121.​1989.tb00368.x

  29. Gao, X., Yu, Q., Yu, R., Brouwers, H., Shui, Z. (2014). Investigation on the effect of slag and limestone powder addition in alkali activated metakaolin. Int. J. Eng. Res. Technol. 3, 123–128. https://doi.org/10.15623/IJRET.2014.0325020

  30. Ghosh, S., Handoo, S. (1980). Infrared and Raman spectral studies in cement and concrete. Cem. Concr. Res. 10, 771–782. https://doi.org/10.1016/0008-8846(80)90005-8

  31. Grubeša, I.N., Barisic, I., Fucic, A., Bansode, S.S. (2016). Characteristics and Uses of Steel Slag in Building Construction. Woodhead Publishing.

  32. Haleatek, T., Opalska, B., Lao, I., Stetkiewicz, J., Rydzynski, K. (2005). Pneumotoxicity of dust from aluminum foundry and pure alumina: A comparative study of morphology and biomarkers in rats. Int. J. Occup. Med. Environ. Health. 18, 59–70.

  33. Hart, J.F., Autenrieth, D.A., Cauda, E., Chubb, L., Spear, T.M., Wock, S., Rosenthal, S. (2018). A comparison of respirable crystalline silica concentration measurements using a direct-on-filter Fourier transform infrared (FT-IR) transmission method vs. a traditional laboratory X-ray diffraction method. J. Occup. Environ. Hyg. 15, 743–754. https://doi.org/10.1080/15459624.​2018.1495334

  34. Hsiao, Y.H., Wang, B., La Plante, E.C., Pignatelli, I., Krishnan, N.A., Le Pape, Y., Neithalath, N., Bauchy, M., Sant, G. (2019). The effect of irradiation on the atomic structure and chemical durability of calcite and dolomite. NPJ Mater. Degrad. 3, 1–9. https://doi.org/10.1038/s41529-019-0098-x

  35. Hughes, T.L., Methven, C.M., Jones, T.G., Pelham, S.E., Fletcher, P., Hall, C. (1995). Determining cement composition by Fourier transform infrared spectroscopy. Adv. Cem. Based Mater. 2, 91–104. https://doi.org/10.1016/1065-7355(94)00031-X

  36. Jadhav, R., Debnath, N. (2011). Computation of X-ray powder diffractograms of cement components and its application to phase analysis and hydration performance of OPC cement. Bull. Mater. Sci. 34, 1137–1150. https://doi.org/10.1007/s12034-011-0134-0

  37. Jiang, X., Wang, S., Ge, L., Lin, F., Lu, Q., Wang, T., Huang, B., Lu, B. (2017). Development of organic–inorganic hybrid beads from sepiolite and cellulose for effective adsorption of malachite green. RSC Adv. 7, 38965–38972. https://doi.org/10.1039/C7RA06351B

  38. Jing, H.P., Li, Y., Wang, X., Zhao, J., Xia, S. (2019). Simultaneous recovery of phosphate, ammonium and humic acid from wastewater using a biochar supported Mg(OH)2/bentonite composite. Environ. Sci. Water Res. Technol. 5, 931–943. https://doi.org/10.1039/C8EW00952J

  39. Jones, J.G., Warner, C. (1972). Chronic exposure to iron oxide, chromium oxide, and nickel oxide fumes of metal dressers in a steelworks. Occup. Environ. Med. 29, 169–177. https://doi.org/​10.1136/oem.29.2.169

  40. Jozić, D., Zelić, J., Janjatović, I. (2010). Influence of the coarse fly ash on the mechanical properties of the cement mortars. Ceram. – Silik. 54, 144–151.

  41. Kasha, A., Al-Hashim, H., Abdallah, W., Taherian, R., Sauerer, B. (2015). Effect of Ca2+, Mg2+ and SO42− ions on the zeta potential of calcite and dolomite particles aged with stearic acid. Colloids Surf., A 482, 290–299. https://doi.org/10.1016/j.colsurfa.2015.05.043

  42. Kasina, M., Kowalski, P.R., Michalik, M. (2015). Mineral carbonation of metallurgical slags. Mineralogia 45, 27–45. https://doi.org/10.1515/mipo-2015-0002

  43. Kicińska, A. (2016). Risk assessment of children's exposure to potentially harmful elements (PHE) in selected urban parks of the Silesian agglomeration. E3S Web. Conf. 10, 00035. https://doi.org/​10.1051/e3sconf/20161000035

  44. Kicińska, A., Bożęcki, P. (2018). Metals and mineral phases of dusts collected in different urban parks of Krakow and their impact on the health of city residents. Environ. Geochem. Health. 40, 473–488, https://doi.org/10.1007/s10653-017-9934-5

  45. Kimata, M. (1980). Crystal chemistry of Ca-melilites on X-ray diffraction and infrared absorption properties. Neues Jahrb. Mineral. 139, 43–58.

  46. Kloprogge, J., Frost, R. (1999). Infrared emission spectroscopy of Al-pillared beidellite. Appl. Clay Sci. 15, 431–445. https://doi.org/10.1016/S0169-1317(99)00025-3

  47. Knaapen, A.M., Borm, P.J.A., Albrecht, C., Schins, R.P.F. (2004). Inhaled particles and lung cancer. Part A: Mechanisms. Int. J. Cancer 109, 799–809. https://doi.org/10.1002/ijc.11708

  48. Kolovos, K., Tsivilis, S., Kakali, G. (2005). SEM examination of clinkers containing foreign elements. Cem. Concr. Compos. 27, 163–170. https://doi.org/10.1016/j.cemconcomp.2004.02.003

  49. Koskinen, K., Pukkala, E., Martikainen, R., Reijula, K., Karjalainen, A. (2002). Different measures of asbestos exposure in estimating risk of lung cancer and mesothelioma among construction workers. J. Occup. Environ. Med. 44, 1190–1196. 

  50. Koval'skii, V., Iarovaia, G.A., Shmavonian, D.M. (1961). Modification of human and animal purine metabolism in conditions of various molybdenum bio-geochemical areas. Zh. Obs. Biol. 22, 179–191. 

  51. Kralevich, M.L., Koenig, J.L. (1998). FTIR analysis of silica-filled natural rubber. Rubber Chem. Technol. 71, 300–309. https://doi.org/10.5254/1.3538486

  52. Kumar, R.S., Rajkumar, P. (2014). Characterization of minerals in air dust particles in the state of Tamilnadu, India through FTIR, XRD and SEM analyses. Infrared Phys. Technol. 67, 30–41. https://doi.org/10.1016/j.infrared.2014.06.002

  53. Lafuente, B., Downs, R.T., Yang, H., Stone, N. (2015). 1. The power of databases: The RRUFF project, in: 1. The power of databases: The RRUFF project, De Gruyter (O), pp. 1–30. https://doi.org/10.1515/9783110417104-003

  54. Laha, T., Balani, K., Agarwal, A., Patil, S., Seal, S. (2005). Synthesis of nanostructured spherical aluminum oxide powders by plasma engineering. Metall. Mater. Trans. A 36, 301–309. https://doi.org/10.1007/s11661-005-0303-0

  55. Li, G., Jing, W., Wen, Z.Q. (2011). Identification of unknown mixtures of materials from biopharmaceutical manufacturing processes by microscopic-FTIR and library searching. Am. Pharm. Rev. 14, 60.

  56. Li, W., Zhou, L., Han, Y., Zhu, Y., Li, Y. (2019). Effect of carboxymethyl starch on fine-grained hematite recovery by high-intensity magnetic separation: Experimental investigation and theoretical analysis. Powder Technol. 343, 270–278. https://doi.org/10.1016/j.powtec.2018.11.024

  57. Manisalidis, I., Stavropoulou, E., Stavropoulos, A., Bezirtzoglou, E. (2020). Environmental and health impacts of air pollution: A review. Front. Public Health 8, 14. https://doi.org/10.3389/​fpubh.2020.00014

  58. Martinez-Tarazona, M.R., Spears, D.A. (1996). The fate of trace elements and bulk minerals in pulverized coal combustion in a power station. Fuel Process. Technol. 47, 79–92. https://doi.org/10.1016/0378-3820(96)01001-6

  59. Maxim, L.D., McConnell, E. (2005). A review of the toxicology and epidemiology of wollastonite. Inhalation Toxicol. 17, 451–466. https://doi.org/10.1080/08958370591002030

  60. Meldrum, M., Howden, P. (2002). Crystalline silica: Variability in fibrogenic potency. Ann. Occup. Hyg. 46, 27–30. https://doi.org/10.1093/annhyg/mef624

  61. Mishra, P.C., Sahu, H.B., Patel, R.K. (2004). Environmental pollution status as a result of limestone and dolomite mining–A case study. Pollut. Res. 23, 427–432. http://hdl.handle.net/2080/157

  62. Meo, S.A. (2004). Health hazards of cement dust. Saudi Med. J. 25, 1153–1159. 

  63. Merget, R., Bauer, T., Küpper, H., Philippou, S., Bauer, H., Breitstadt, R., Bruening, T. (2002). Health hazards due to the inhalation of amorphous silica. Arch. Toxicol. 75, 625–634. https://doi.org/​10.1007/s002040100266

  64. Mohiuddin, K., Strezov, V., Nelson, P.F., Stelcer, E., Evans, T. (2014). Mass and elemental distributions of atmospheric particles nearby blast furnace and electric arc furnace operated industrial areas in Australia. Sci. Total Environ. 487, 323-334. https://doi.org/10.1016/j.​scitotenv.2014.04.025

  65. Mollah, M.Y.A., Tsai, Y.N., Cocke, D.L. (1992). An FTIR investigation of cement based solidification/stabilization systems doped with cadmium. J. Environ. Sci. Health A 27, 1213–1227. https://doi.org/10.1080/10934529209375792

  66. Mueller, E.J., Seger, D.L. (1985). Metal fume fever—A review. J. Emergency Med. 2, 271–274. https://doi.org/10.1016/0736-4679(85)90106-4

  67. Nagabhushana, H., Prashantha, S., Nagabhushana, B., Lakshminarasappa, B., Singh, F. (2008). Damage creation in swift heavy ion-irradiated calcite single crystals: Raman and infrared study. Spectrochim. Acta, Part A 71, 1070–1073. https://doi.org/10.1016/j.saa.2008.03.003

  68. Neghab, M. (2015). Respiratory morbidity associated with long-term occupational inhalation exposure to high concentrations of hydrated calcium sulfate dust. Occup. Diseases Environ. Med. 4, 1. https://doi.org/10.4236/odem.2016.41001

  69. Neupane, B.B., Sharma, A., Giri, B., Joshi, M.K. (2020). Characterization of airborne dust samples collected from core areas of Kathmandu Valley. Heliyon 6, 03791. https://doi.org/10.1016/​j.heliyon.2020.e03791

  70. Oberdörster, G. (1993). Lung dosimetry: Pulmonary clearance of inhaled particles. Aerosol Sci. Technol. 18, 279–289. https://doi.org/10.1080/02786829308959605

  71. Ojima, J. (2003). Determining of crystalline silica in respirable dust samples by infrared spectrophotometry in the presence of interferences. J. Occup. Health. 45, 94–103. https://doi.org/10.1539/joh.45.94

  72. Pandharkar, S.M., Rondiya, S.R., Rokade, A.V., Gabhale, B.B., Pathan, H.M., Jadkar, S.R. (2018). Synthesis and characterization of molybdenum back contact using direct current-magnetron sputtering for thin film solar cells. Front. Mater. 5, 13. https://doi.org/.3389/fmats.2018.00013

  73. Parada, F., Parra, R., Watanabe, T., Hino, M., Palacios, J., Sánchez, M. (2009). Proc. VIII Int. Conf. on ‘Molten slags, fluxes and salts’, Santiago, Chile, 2009.

  74. Park, S., Wexler, A. (2008). Size-dependent deposition of particles in the human lung at steady-state breathing. J. Aerosol Sci. 39, 266–276. https://doi.org/10.1016/j.jaerosci.2007.11.006

  75. Rafeet, A., Vinai, R., Soutsos, M., Sha, W. (2019). Effects of slag substitution on physical and mechanical properties of fly ash-based alkali activated binders (AABS). Cem. Concr. Res. 122, 118–135. https://doi.org/10.1016/j.cemconres.2019.05.003

  76. Rautray, T., Behera, B., Badapanda, T., Vijayan, V., Panigrahi, S. (2009). Trace element analysis of fly ash samples by EDXRF technique. Indian J. Phys. 83, 543–546. https://doi.org/10.1007/​s12648-009-0016-0

  77. Reig, F.B., Adelantado, J.G., Moreno, M.M. (2002). FTIR quantitative analysis of calcium carbonate (calcite) and silica (quartz) mixtures using the constant ratio method. Application to geological samples. Talanta 58, 811–821. https://doi.org/10.1016/S0039-9140(02)00372-7

  78. Sandberg, W.J., Låg, M., Holme, J.A., Friede, B., Gualtieri, M., Kruszewski, M., Schwarze, P.E., Skuland, T., Refsnes, M. (2012). Comparison of non-crystalline silica nanoparticles in Il-1β release from macrophages. Part. Fibre Toxicol. 9, 32. https://doi.org/10.1186/1743-8977-9-32

  79. Sedaghat, A., Ram, M.K., Zayed, A., Kamal, R., Shanahan, N. (2014). Investigation of physical properties of graphene-cement composite for structural applications. Open J. Compos. Mater. 2014. https://doi.org/10.4236/ojcm.2014.41002

  80. Seguin, L., Figlarz, M., Cavagnat, R., Lassègues, J.C. (1995). Infrared and Raman spectra of MoO3 molybdenum trioxides and MoO3· XH2O molybdenum trioxide hydrates. Spectrochim. Acta, Part A 51, 1323–1344. https://doi.org/10.1016/0584-8539(94)00247-9

  81. Shao, Y., El-Baghdadi, A., He, Z., Mucci, A., Fournier, B. (2015). Carbon dioxide–activated steel slag for slag-bonded wallboard application. J. Mater. Civ. Eng. 27, 04014119. https://doi.org/​10.1061/(ASCE)MT.1943-5533.0001055

  82. Sharma, S.K., Simons, B., Yoder, H. (1983). Raman study of anorthite, calcium Tschermak's pyroxene, and gehlenite in crystalline and glassy states. Am. Mineral. 68, 1113–1125. 

  83. Shih, W.Y., Rahardianto, A., Lee, R.W., Cohen, Y. (2005). Morphometric characterization of calcium sulfate dihydrate (gypsum) scale on reverse osmosis membranes. J. Membr. Sci. 252, 253–263. https://doi.org/10.1016/j.memsci.2004.12.023

  84. Snellings, R., Mertens, G., Elsen, J. (2012). Supplementary cementitious materials. Rev. Mineral. Geochem. 74, 211–278. https://doi.org/10.2138/rmg.2012.74.6

  85. Sofilić, T., Mladenović, A., Oreščanin, V., Barišić, D. (2013). In 13th International Foundrymen Conference, Inovative Foundy Processe and Materials, F. Unkić (ur.), Opatija, 2013, pp. 354–369.

  86. Soheilmoghaddam, M., Wahit, M.U., Yussuf, A.A., Al-Saleh, M.A., Whye, W.T. (2014). Characterization of bio regenerated cellulose/sepiolite nanocomposite films prepared via ionic liquid. Polym. Test. 33, 121–130. https://doi.org/10.1016/j.polymertesting.2013.11.011

  87. Stutzman, P.E., Feng, P., Bullard, J.W. (2016). Phase analysis of portland cement by combined quantitative X-ray powder diffraction and scanning electron microscopy. J. Res. Natl. Inst. Stand. Technol. 121, 47–107. https://doi.org/10.6028/jres.121.004

  88. Suryadevara, B. (2016). Advances in Chemical Mechanical Planarization (Cmp). Woodhead Publishing.

  89. Taylor, W. (1990). Application of infrared spectroscopy to studies of silicate glass structure: Examples from the melilite glasses and the systems Na2O-SiO2 and Na2O-Al2O3-SiO2. Proc. Indian Acad. Sci. Math. Sci. 99, 99–117. https://doi.org/10.1007/BF02871899

  90. Tazaki, K., Fyfe, W., Sahu, K., Powell, M. (1989). Observations on the nature of fly ash particles. Fuel 68, 727–734. https://doi.org/10.1016/0016-2361(89)90211-1

  91. Tiwari, M., Sahu, S., Bhangare, R., Ajmal, P., Pandit, G. (2014). Elemental characterization of coal, fly ash, and bottom ash using an energy dispersive X-ray fluorescence technique. Appl. Radiat. Isot. 90, 53–57. https://doi.org/10.1016/j.apradiso.2014.03.002

  92. Toffolo, M.B., Regev, L., Dubernet, S., Lefrais, Y., Boaretto, E. (2019). FTIR-based crystallinity assessment of aragonite–calcite mixtures in archaeological lime binders altered by diagenesis. Minerals 9, 121. https://doi.org/10.3390/min9020121

  93. Tolinggi, S., Nakoe, M.R., Gobel, I.A., Sengke, J., Keman, S., Sudiana, I.K., Yudhastuti, R., Azizah, R. (2014). Effect inhaling of limestone dust exposure on increased level of IL-8 serum and pulmonary function decline to workers of limestone mining industry. Int. J. Eng. Sci. 3, 66–72. https://repository.unair.ac.id/105374/

  94. Tripathy, S.K., Dasu, J., Murthy, Y.R., Kapure, G., Pal, A.R., Filippov, L.O. (2020). Utilisation perspective on water quenched and air-cooled blast furnace slags. J. Cleaner Prod. 262, 121354. https://doi.org/10.1016/j.jclepro.2020.121354

  95. Valiulis, D., Šakalys, J., Plauškaitė, K. (2008). Heavy metal penetration into the human respiratory tract in Vilnius. Lith. J. Phys 48, 349–355. https://doi.org/10.3952/lithjphys.48407

  96. Wang, K., Ren, L., Yang, L. (2018). Excellent carbonation behavior of rankinite prepared by calcining the C-S-H: Potential recycling of waste concrete powders for prefabricated building products. Materials 11, 1474. https://doi.org/10.3390/ma11081474

  97. Wang, X., Zhang, Y., Lv, L., Cui, Y., Wei, C., Pang, G. (2016). Preparation of Mg(Oh)2 hybrid pigment by direct precipitation and graft onto cellulose fiber via surface-initiated atom transfer radical polymerization. Appl. Surf. Sci. 363, 189–196. https://doi.org/10.1016/j.apsusc.2015.11.259

  98. Wilczyńska-Michalik, W., Rzeźnikiewicz, K., Pietras, B., Michalik, M.(2015). Fine and ultrafine TiO2 particles in aerosol in Kraków (Poland). Mineralogia 45, 65–77. https://doi.org/10.1515/mipo-2015-0005

  99. Wójcik, E., Szwajgier, P., Oleszczuk, A., Mieczan, W. (2020). Effects of titanium dioxide nanoparticles exposure on human health-A review. Biol. Trace Elem. Res. 193, 118–129. https://doi.org/10.1007/s12011-019-01706-6

  100. Wyers, H. (1949). Asbestosis. postgrad. Med. J. 25, 631. https://doi.org/10.1136/pgmj.25.290.631

  101. Yang, C.Y., Huang, C.C., Chiu, H.F., Chiu, J.F., Lan, S.J., Ko, Y.C. (1996). Effects of occupational dust exposure on the respiratory health of Portland cement workers. J. Toxicol. Environ. Health 49, 581–588.

  102. Yehia, H.M., El-Tantawy, A., Ghayad, I., Eldesoky, A.S., El-kady, O. (2020). Effect of zirconia content and sintering temperature on the density, microstructure, corrosion, and biocompatibility of the Ti–12Mo matrix for dental applications. J. Mater. Res. Technol. 9, 8820–8833. https://doi.org/10.1016/j.jmrt.2020.05.109

  103. Yip, C.K., Provis, J.L., Lukey, G.C., van Deventer, J.S. (2008). Carbonate mineral addition to metakaolin-based geopolymers. Cem. Concr. Compos. 30, 979–985. https://doi.org/10.1016/j​.cemconcomp.2008.07.004

  104. Yusan, S., Gok, C., Erenturk, S., Aytas, S. (2012). Adsorptive removal of thorium (IV) using calcined and flux calcined diatomite from Turkey: Evaluation of equilibrium, kinetic and thermodynamic data. Appl. Clay Sci. 67, 106–116. https://doi.org/10.1016/j.clay.2012.05.012

  105. Zhang, M., Chen, T., Wang, Y. (2017). Insights into TiO2 polymorphs: Highly selective synthesis, phase transition, and their polymorph-dependent properties. RSC Adv. 7, 52755–52761. https://doi.org/10.1039/C7RA11515F

  106. Zhang, S., Worrell, E., Crijns-Graus, W. (2015). Evaluating co-benefits of energy efficiency and air pollution abatement in China’s cement industry. Appl. Energy 147, 192–213. https://doi.org/​10.1016/j.apenergy.2015.02.081

  107. Zhang, W., Lv, C. (2020). Effects of mineral content on limestone properties with exposure to different temperatures. J. Pet. Sci. Eng. 188, 106941. https://doi.org/10.1016/j.petrol.2020.​106941

  108. Zhang, W., Zhang, S., Wang, J., Dong, J., Cheng, B., Xu, L., Shan, A. (2019). A novel adsorbent albite modified with cetylpyridinium chloride for efficient removal of zearalenone. Toxins. 11, 674. https://doi.org/10.3390/toxins11110674

  109. Zhang, Z., Wang, H., Provis, J.L. (2012). Quantitative study of the reactivity of fly ash in geopolymerization by FTIR. J. Sustainable Cem.-Based Mater. 1, 154–166. https://doi.org/​10.1080/21650373.2012.752620

  110. Zong, Y., Wan, Q., Cang, D. (2019). Preparation of anorthite-based porous ceramics using high-alumina fly ash microbeads and steel slag. Ceram. Int. 45, 22445–22451. https://doi.org/​10.1016/j.ceramint.2019.08.003

  111. Żyrkowski, M., Neto, R.C., Santos, L.F., Witkowski, K. (2016). Characterization of fly-ash cenospheres from coal-fired power plant unit. Fuel 174, 49–53. https://doi.org/10.1016/​j.fuel.2016.01.061


Share this article with your colleagues 

 

Subscribe to our Newsletter 

Aerosol and Air Quality Research has published over 2,000 peer-reviewed articles. Enter your email address to receive latest updates and research articles to your inbox every second week.

7.3
2022CiteScore
 
 
77st percentile
Powered by
Scopus
 
   SCImago Journal & Country Rank

2021 Impact Factor: 4.53
5-Year Impact Factor: 3.668

The Future Environment and Role of Multiple Air Pollutants

Aerosol and Air Quality Research partners with Publons

CLOCKSS system has permission to ingest, preserve, and serve this Archival Unit
CLOCKSS system has permission to ingest, preserve, and serve this Archival Unit

Aerosol and Air Quality Research (AAQR) is an independently-run non-profit journal that promotes submissions of high-quality research and strives to be one of the leading aerosol and air quality open-access journals in the world. We use cookies on this website to personalize content to improve your user experience and analyze our traffic. By using this site you agree to its use of cookies.