Next Article in Journal
Weathered Sand of Basalt as a Potential Nickel Adsorbent
Next Article in Special Issue
Tribological Properties of Additive Manufactured Materials for Energy Applications: A Review
Previous Article in Journal
The Hydrodynamics and Mixing Performance in a Moving Baffle Oscillatory Baffled Reactor through Computational Fluid Dynamics (CFD)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Behavior of Cd during Coal Combustion: An Overview

ENET Centre, VSB-Technical University of Ostrava, 708 00 Ostrava-Poruba, Czech Republic
*
Author to whom correspondence should be addressed.
Processes 2020, 8(10), 1237; https://doi.org/10.3390/pr8101237
Submission received: 3 September 2020 / Revised: 25 September 2020 / Accepted: 29 September 2020 / Published: 2 October 2020
(This article belongs to the Special Issue Recent Advances in Processed Materials for Energy Applications)

Abstract

:
Due to the unfavorable combination of its toxicity and high volatility, Cd is contained in most lists of potentially hazardous air pollutants with the greatest environmental and human-health concerns. The review paper evaluates the behavior of Cd during combustion (incineration) processes and its redistribution among condensed fractions (bottom ash/slag, fly ash) and volatilized fractions (that passes through most particulate control devices). The paper addresses all important effects of Cd interactions, such as presence of organic or inorganic chlorides, moisture levels, S, P and Na concentrations, flue gas composition etc. Possibilities of using various adsorbents (either within in-furnace regime or applied in post-combustion zone) are evaluated as well. Special attention is paid to mitigating its emissions factors; decreasing Cd volatility and facilitating Cd retention are discussed with the view of various combustion (incineration) conditions and the feed fuel composition.

1. Introduction

Coal contains almost all the elements from the periodic chart. These elements differ by their concentrations, modes of occurrence, toxicity and behavior during coal combustion at power stations [1]. These factors influence not only the perspectives of their potential extractions from coal/ashes (such as rare earth elements (REE) [2,3], Y [4], Ge-Ga [5] etc.) but also their negative impact on the environment.
Environmental hazards are created by the coal characteristics (including the levels and associations of interacting species) and additionally by the operational conditions, such as the combustion technology, temperature, atmosphere, air-pollution control devices and proper use of desulfurization additives [6,7,8] or adsorbents [9,10,11].
Despite the progress achieved during the last decades, the decreasing of emissions of toxic elements is still a tough challenge. The overall environmental and human-health hazards of individual elements are given by their partitioning among bottom ash (slag), fly ash and emissions, as well as by their toxicity.
Wagner and Hlatshwayo [12] summarized and evaluated six lists of potentially hazardous air pollutants. All these lists contain Hg and As, two toxic elements exhibiting high volatility, which is in line with wide attention in the literature that is paid to Hg [13,14,15] and As [8,16,17]. Another two elements are present in all these lists—Pb and Cd. The behavior of and possibilities of decreasing Pb emissions were discussed in our previous paper [18]; therefore, this review paper is focused on Cd, due to the unfavorable combination of its high toxicity and volatile character hindering its quantitative retention.
Owing to the increasing amount of industrial, agricultural, forest and domestic wastes, there is a trend toward co-combustion of coal with these waste materials. For example, the amount of municipal solid waste (MSW) of the cities (in world-wide scale) may reach 2.2 billion tonnes per year by 2025 [19,20]. Moreover, if an effective and efficient solid waste management is not applied, it can result in health hazards or negative impact on the environment [21]. Therefore, possible interactions of various wastes with coal during co-combustion is discussed herein as well.
The objective of this review paper is to address all important aspects relating to Cd behavior and interactions during combustion (incineration) processes with particular attention paid on decreasing its emissions.

2. Cd (and Cl) in Coal

2.1. Concentrations of Cd

Clarke values for Cd concentrations in low-rank and high-rank coals are 0.24 and 0.20 ppm, which is ca. one- or two-orders of magnitude lower than those of Pb and Cr [22]. Since Cd concentrations in coal are rather variable, Biswas and Wu [23] provided typical Cd concentrations range in coal (entering combustors) being 0.033–0.64 ppm. For the comparison, Table 1 provides also Cd levels in other related materials, such as in municipal solid waste (ash) [24] or some coal combustion additives (limestone or urea) [25]. It can be deduced from these data that if some of these materials are co-combusted with coal (e.g., in fluidized-bed power stations) the amount of Cd introduced into the combustion chamber with a waste might be significantly higher than its input mass in coal. In contrast, if limestone (ca. 0.01 ppm Cd) or urea (ca. 0.03 ppm Cd) [25] are put into the furnace due to flue gas cleaning, their low Cd concentrations might result in a diluting effect.

2.2. Associations of Cd

The sulfidic (chalcophylic) associations of Cd [26,27] is even more dominant than that of Pb (ca. 80% Cd in low-rank coals and 65% Cd in high-rank coals [28]), which relates primarily to Cd occurrence in sphalerite. There is also some affinity to pyrite (mainly in high rank coals—25%) and ca. 10% silicate-associated Cd in both high-rank and low-rank coals [28]. Organic affinity of Cd has been observed in Waterberg Coalfield coal [29], Huaibei Coalfield coal [26] or in coal form Upper Silesian Coal basin [30].
During coal combustion, Cd behaves as one of the most volatile elements, which corresponds well with its dominant sulfidic affinity accompanied by minor occurrence in organic matter or carbonates; according to Finkelman [28], only ca. 10% of Cd (in average) is firmly bound in (alumino) silicates.

2.3. Cl in Coal (and Wastes)

Due to the abundant and widely-discussed interactions of Cd with Cl during combustion, Cl levels in coal are important as well. Clarke values of estimated Cl concentrations in brown coals and hard coals (and their ashes) are 120 and 340 ppm (and 770 and 2100 ppm in corresponding ashes) [31,32]. Therefore, even if Cd concentrations in low-rank and high-rank coals are similar, higher Cl content in high-rank coals might favor Cd volatilization (release) from these coals.
Even though organically-bound (covalent bond) Cl may also be present in coals, its occurrence in inorganic chlorides (NaCl, KCl etc.) is undoubtedly the most abundant; some “semi-organic” Cl (i.e., anion Cl sorbed on organic matter from pore water) was observed as well [31,32].
Fluidized-bed combustion technology is spreading quickly around the world [33] which enables to use also low-grade fuels (biomass, wastes, etc.) whose composition and Cl levels are of fluctuating quality. For example, refuse-derived fuel typically contains PVC (polyvinyl chloride), sewage sludge often contains higher levels of FeCl3, NaCl is present in food residues and biomass may contain higher Cl (and moisture) concentrations [34,35,36]. All these different Cl forms exhibit different effects on Cd volatilities (which will be discussed in Section 5).

3. Melting/Boiling Points of Cd Compounds

Based on thermodynamic equilibrium calculations in oxidizing atmosphere (not considering Cd-Si and Cd-Al interactions) [37], the most abundant Cd-compounds during coal combustion are CdO, CdCl2, CdSO4 and Cd. Melting/boiling points of these compounds are crucial for detailed evaluation of Cd transformations during coal combustion and are listed in Table 2.
It is noteworthy to mention also the melting points of some the most abundant interacting species that typically occur during coal combustion, e.g., NaCl (801 °C) [42], KCl (770 °C) [42], CaCl2 (ca. 800 °C) [43], CaSO4 (1400 °C) [43].
As melting/boiling points of CdCl2 are lower than those of CdO or CdSO4, formation of CdCl2 leads to increased volatility of Cd, which is quite hindering to its quantitative capture (which will be discussed in detail in Section 5). The fraction of Cd that does not condense during flue gas cooling or is not retained by additives or ash particles might remain in gaseous form that can easily pass through particulate control device (electrostatic precipitator or fabric filter) as depicted in Figure 1.

4. Combustion/Retention Experiments without Extra Cl Added

4.1. In-Furnace Adsorbents

Various adsorbents were tested for Cd retention in combustion/incineration processes (bauxite, kaolinite, Al2O3, silica sand, apatite, zeolite, mullite, scallop etc.) Nevertheless, as efficient Cd retention is a complex phenomenon (depending at least on the temperature, dwell time, atmosphere and concentrations of interacting species) [44], results observed in individual combustion/retention experiments are quite variable, which is shown in Table 3. For example, bauxite is a promising adsorbent of Cd. Nevertheless, in different studies [45,46,47,48,49] the retained fraction of Cd is rather fluctuating—from 7–14% [46,48,49] up to 68–74% [45,47] despite the combustion temperature was similar (700–800 °C), which can be attributed to different composition of the feed and different experimental conditions (including time factor). Similar results are documented also for kaolin(ite) that is widely used for Cd retention as well [45,46,48,49,50,51]—herein, the retention efficiency ranges from 4% [46] up to 44% [48,49].

4.1.1. Bauxite

An effective retention of Cd can be achieved by bauxite [45,46,47], mostly at lower temperatures (700–800 °C) [45,47]. However, at higher temperatures (1100 °C) the ability of bauxite to retain Cd was observed as well [50]—in this experiment, adsorbent pre-mixed with coal was combusted in a vertical tube reactor at 1100 °C and bauxite provided better retention results than kaolinite and CaO (based on evaluation of particles with medium size). At lower temperatures (700 or 900 °C), during synthetic MSW incineration bauxite was better adsorbent of Cd than kaolinite and Al2O3 [46], which might indicate that the admixture of other minerals/components in bauxite might favor Cd retention in comparison with pure Al2O3. Moreover, (at 700 °C) bauxite was the best adsorbent not only for Cd but also for Cr, Pb and Cu. Interaction of CdO and CdCl2 with Al2O3 can be described by Reactions (1) [47] and (2) [45,47].
C d O + A l 2 O 3 C d A l 2 O 4 ( s )
C d C l 2 + A l 2 O 3 + H 2 O C d O · A l 2 O 3 ( s ) + 2 H C l ( g )

4.1.2. Kaolinite

Yao and Naruse [48,49] tested 9 adsorbents for in-furnace retention of Cd during dried sewage sludge combustion in drop-tube furnace at 800 °C. The most efficient Cd capture was observed in the case of kaolinite (31.55 m2/g), followed by zeolite (21.40 m2/g), limestone (1.87 m2/g), scallop (0.85 m2/g), apatite (35.14 m2/g) ~mullite (8.60 m2/g), bauxite (0.92 m2/g), silica (1.29 m2/g) and alumina (1.23 m2/g)—captured fractions of Cd are given in Table 3. The results indicate that specific surface area is not the only affecting factor. Efficient retention by kaolinite can be described by Reaction (3) [48,49] and is in line with the conclusion that the retained Cd in residual ash may be in the form of binary oxides C d O · A l 2 O 3 or C d O · S i O 2   [25].
C d C l 2 + k a o l i n ( A l 2 O 3 · 2 S i O 2 · 2 H 2 O ) C d O · A l 2 O 3 · 2 S i O 2 + 2 H C l
Kaolinite was also tested during high-temperature combustion. Chen et al. [51] presented better retention of Cd on kaolin than on CaO if added to sludge prior to combustion at 1200 °C (in TGA); however, it is interesting that absolutely the best retention results were observed if no adsorbent was added to the sludge (as shown in Table 3).
Wendt and Lee [54] document efficient retention of Cd (and Pb) using blended adsorbent containing kaolinite and calcite (with admixture of lime) injected into flame in vertical combustor at 1160 and 1280 °C; the former temperature was better for the retention of Pb, the latter was advantageous for the capture of Cd (if there was no Pb). If both Cd and Pb were present, the lower temperature (1160 °C) was good also for the retention of Cd, which was interpreted by surface melt formed on the adsorbent particles due to the enhancing effect of Pb.

4.1.3. Silica Sand

Silica sand can also adsorb heavy metals (HMs) including Cd [55,56]. However, in both studies, the retention of Cd was somewhat worse than that of Pb and can be described by Reaction (4) [47]
C d O + S i O 2 C d S i O 3 ( s )
The effect of Na (if NaNO3 was added to silica sand) on the retention of Cd and Pb was studied by Kuo et al. [53]. The presence of Na increased the retention ability of silica sand at 700–900 °C; at 900 °C the increase was the most significant—from ca. 5% retained Cd (no Na) up to ca. 46% (for 1.2% Na). If Na was added at 800 and 900 °C, the retention of Cd was even better than that of Pb (ca. 41% Pb vs. 46% Cd at 900 °C). The observation was attributed to the formation of low-melting-point eutectics on the adsorbent particle surfaces; the effect of physical co-condensation was presumed as being also possible.

4.1.4. Calcareous Adsorbents

Calcareous adsorbents were tested for the retention of Cd as well. In general, if no extra Cl (or S) was added to the feed, then bauxite, kaolinite or zeolite typically provide more efficient retention (than calcareous adsorbents). However, in the case of higher Cl (or S) levels in the feed, calcareous adsorbents (CaCO3 or CaO) provided usually better results, which will be discussed in detail in Section 5.2.1.

4.2. Other Approaches

Instead of adding kaolin directly into the furnace, suspension of water with kaolinite can be used also in semi-dry spray tower at 150–170 °C [57]. Synthetic solid waste was combusted at 800 °C and in semi-dry spray tower limestone, kaolinite, Al2O3 or no sorbent (i.e., only water) were applied. For the retention of Cd, the best results were achieved by spraying with water without any adsorbent, while in the case of other HMs, added adsorbents provided better results than water.
Peng et al. [55] reported promising Cd-retention results by means of low-temperature two-stage fluidized bed incinerator where artificial MSW was combusted. The temperature at the first stage was 550, 650 and 750 °C, at the second (filtration) stage it was always 800 °C. The best Cd retention was obtained at 550 °C (at the first stage). Efficient Cd capture in this low-temperature two-stage system was attributed to Cd vaporization, chemical adsorption and filtration in the second stage, which is in line with conclusions of other studies [58,59]. It is worth mentioning that Cd retention at 650 °C (the 1st stage) was lower than that at 550 °C but almost all this decrease was retained later during the filtration at the 2nd stage. Then, the overall retention of Cd at 650 °C was only slightly lower than that at 550 °C (due to efficient filtration at the second stage).

5. Effect of Cl

In general, the effect of Cl on behavior of Cd (and other HMs) is studied namely due to shifting CdO-CdCl2 equilibrium toward the formation of chloride [60]. As chlorides of most metals (including Cd) exhibit higher volatility than oxides, there are two dominant reasons why attention is paid to this phenomenon:
(i)
Due to enhanced emissions of Cd (and other HMs) during combustion/incineration processes (high volatility of CdCl2 favors Cd occurrence in gaseous form that might easily pass through the air-pollution control device).
(ii)
Due to removal of HMs from combustion (incineration) ashes which facilitates their further technological utilization. High volatility of CdCl2 facilitates Cd release from bottom ash leading to lower Cd concentrations and better ash utilization perspectives. In contrast, due to volatilization/condensation mechanism, certain fraction of Cd might condense/adsorb on fly ash particles increasing Cd levels there (Figure 1). Moreover, these fractions often exhibit increased leachability.
Affinity of individual elements to Cl is different—approximate sequences of these affinities can be found in the literature [61,62,63,64]:
T l > C u > Z n > P b > C o > M n > S n > H g
C u > T l > ( S n , Z n ) > ( P b , C d ,   N i , C o ,   S b )
H > N a > o t h e r   H M s
H > ( N a , K ) > P b
There is a consensus in the literature that Cd exhibits lower affinity to Cl than Pb [65].
The idea of close partitioning of Cd and Cl during coal combustion is supported (i.a.) by strong positive linear relationship between relative enrichment factors of Cd and Cl in filter ash and cyclone ash [66].
Papers documenting the (typically enhancing) effect of Cl on Cd volatility are abundant in the literature—Table 4 provides at least some typical illustrative examples of this effect. As shown in Table 4, various Cl-bearing compounds are able to influence Cd partitioning, such as NaCl, PVC, NH4Cl, FeCl3, CaCl2 or MgCl2. Despite some similar features, the effect of individual chlorination agents is different.
There is a consensus in the literature that there are two general chlorination mechanisms—direct and indirect one. Direct chlorination is thought to be a prevalent mechanism in the case of NaCl where (high-temperature) direct reaction between NaCl and Cd (or other HMs) oxides occur [67]. Indirect chlorination is more common and was observed (e.g.,) in the case of PVC, CaCl2 or MgCl2 and is based on (low-temperature) release of chlorination agent (HCl, Cl2) that is followed by its reaction with CdO (or other HMs oxides) [68].

5.1. Direct Chlorination (Effect of NaCl)

Direct chlorination mechanism of Cd (and other HMs) is typically discussed in the case of NaCl [71], which is present in coals, MSW (food residues), sludges, calcareous additives etc.
During simulated MSW incineration in horizontal tube furnace (900 °C, 60 min), NaCl did not lowered volatilization temperature of Cd (it remains within the range 600 °C–700 °C) but NaCl enhanced the volatilization rates at 700 °C–900 °C [68]. Enhanced Cd volatilization initiated at 700 °C–800 °C, which is lower than the temperature of HCl release from NaCl (t > 900 °C); therefore, the indirect chlorination mechanism is less probable in this case [68].
Based on thermodynamic calculations, it was concluded that both SiO2 and Al2O3 are needed for direct chlorination via NaCl (Reaction (5)) [68]:
2 N a C l + C d O + 2 S i O 2 ( c ) + A l 2 O 3 ( c ) C d C l 2 ( g ) + 2 N a A l S i O 4 ( c )
According to Nowak et al. [71] and Chan et al. [41], reaction of NaCl with H2O and O2 (releasing HCl and Cl2) is less probable than evaporation of NaCl without reacting. However, in the case of direct chlorination, NaCl evaporation or melting is favorable because the mixture of Cl and Cd (or other HM) is not typically homogeneous.
This is consistent with the conclusion of Yu et al. [42]: when temperature reaches melting point of NaCl (801 °C) or KCl (770 °C), NaCl or KCl become liquid phase; then, the reaction rates between liquid NaCl/KCl and SiO2 or Al2O3 can be accelerated. At higher temperatures, NaCl or KCl was further vaporized into gas phase and the formation of HCl or Cl2 through gas-solid reaction can occur (as described in Reactions (2)–(6)) [42,67]:
2 N a C l + S i O 2 + O 2 N a 2 O · S i O 2 + C l 2
2 N a C l + S i O 2 + H 2 O N a 2 O · S i O 2 + 2 H C l
2 N a C l + 2 S i O 2 + A l 2 O 3 + H 2 O 2 N a A l S i O 4 + 2 H C l
2 N a C l + 6 S i O 2 + A l 2 O 3 + H 2 O 2 N a A l S i 3 O 8 + 2 H C l
2 N a C l + 2 S i O 2 + H 2 O N a 2 S i 2 O 5 + 2 H C l
In contrast, CaCl2 or MgCl2 directly react with H2O and/or O2 releasing HCl and/or Cl2 (i.e., Al2O3 or SiO2 are not needed for this release) and are thought to be more efficient chlorination agents.
However, it should be mentioned in this context that when MSW incineration fly ash was roasted in the rotary reactor at 1000 °C (60 min), added NaCl even slightly lowered the Cd removal [71]. The authors hypothesized that NaCl could probably form azeotropes (vapor pressure minima) with HM chlorides.
The results are opposite to those of Wang et al. [68] (where NaCl enhanced Cd volatilization at 700 °C–900 °C); probably due to higher temperature (1000 °C) used in the study of Nowak et al. [71] where NaCl might favor surface melt on the particles which supports Cd retention and thereby hinders Cd evaporation.

5.2. Indirect Chlorination

Indirect chlorination is more common than direct mechanism and is based on release of HCl or Cl2 from Cl-donator followed by interaction of Cl-containing species (HCl, Cl2) with CdO (or other HM oxides) typically increasing its volatility.
Step 1 is the formation of HCl/Cl2 by interaction with O2/H2O [24,32,42] (Reactions (7)–(9)):
M g C l 2 ( o r   C a C l 2 ) + 1 2 O 2 M g O ( o r   C a O ) + C l 2
M g C l 2 ( o r   C a C l 2 ) + H 2 O M g O ( o r   C a O ) + 2 H C l
C l 2 + H 2 O 2 H C l + 1 2 O 2
Interactions with (alumino)silicates are given [24] by Reactions (10)–(13):
C a C l 2 ( l ) + H 2 O ( g ) + S i O 2 ( s ) C a S i O 3 ( s ) + 2 H C l ( g )
C a C l 2 ( l ) + 1 2 O 2 ( g ) + S i O 2 ( s ) C a S i O 3 ( s ) + C l 2 ( g )
C a C l 2 ( l ) + H 2 O ( g ) + A l 2 O 3 ( s ) C a A l 2 O 4 ( s ) + 2 H C l ( g )
C a C l 2 ( l ) + 1 2 O 2 ( g ) + A l 2 O 3 ( s ) C a A l 2 O 4 ( s ) + C l 2 ( g )
Step 2 is chlorination of CdO (in oxidizing atmosphere) [72] (Reaction (14)):
C d O ( s ) + 2 H X ( g ) C d X 2 ( g ) + H 2 O ( g )
where X = Cl, Br, F, I.
Since Cl levels are typically higher than those of Br, F and I, interactions of CdO and HCl are expected to be more abundant than those with HBr, HF and HI.
Individual chlorination agents exhibit different tendency to release Cl, which can be quantified by equilibrium partial pressures over these chlorination agents; equilibrium partial pressures increase in the sequence [41]:
N a C l > C a C l 2 > F e C l 2 > M g C l 2 > A l C l 3   ( 600 1200   ° C )
Cd removal during muffle-furnace incineration tests with MSW incineration ashes was evaluated after addition of NaCl, MgCl2 and CaCl2 in the temperature range of 800 °C–1200 °C [71]. Due to the naturally high volatility of Cd, the experiments at 900–1200 °C were quite inconclusive for all chlorination agents as all removal efficiencies (volatilities) were nearly 100%. Nevertheless, at 800 °C, only increasing amount of MgCl2 and CaCl2 enhanced original Cd volatility (50%) up to ca. 90% in the case of MgCl2 (100–150 g Cl/kg) or up to ca. 85% in the case of CaCl2 (150–200 g Cl/kg). As these experiments were conducted with the aim to remove Cd (and other HMs) from the ash, roasting time was set as 20 h. Therefore, application of these results for the estimation of the volatility of Cd during e.g., fluidized-bed combustion (FBC) at 850 °C can be quite misleading as dwell time in high-temperature zone in fluidized bed is significantly shorter.
These results are consistent with very high volatility of Cd at 1000 °C (60 min) when addition of NaCl, AlCl3, MgCl2, FeCl3 and CaCl2 were tested. Without these agents, the Cd removal exceeded 80% and if 0.3 g Cl was added (by aforementioned agents) the removal reached ca. 95% and was comparable for all these chlorination agents [41]. Nevertheless, also in this case, 60 min roasting time is still too long to be used also for the evaluation of Cd volatility at power station (these experiments were conducted with the aim to remove HMs from the ash).
Indirect chlorination mechanism was considered as the most probable underlying mechanism of MgCl2 [71]; in the case of CaCl2 it prevails as well (even if direct chlorination might also occur [71]).

5.2.1. Indirect Chlorination by PVC and Interactions with Calcareous Minerals

A typical example of an indirect chlorination agent is PVC releasing HCl at ca. 240 °C (in oxidizing atmosphere) [68,73,74], which corresponds with low-temperature volatilization.
During incineration experiments with simulated MSW [68], PVC lowered initial volatilization temperature from 600–700 °C to ca. 500 °C (which is initial volatilization temperature of CdCl2). At higher temperatures (t > 600 °C) it enhanced the vaporization rates (in contrast to NaCl that did not lower volatilization temperature). In the presence of PVC in the feed waste, Cd reached a maximum volatilization rate at 600–700 °C, which is lower than in the case of Pb (800 °C–900 °C).
HCl released from PVC (or other indirect chlorination agents) enhances the volatility of Cd and other HMs, unless it is retained by ash components, e.g., by calcareous additives used during FBC [75,76,77].
There is a consensus in the literature that CaO can adsorb HCl (g) forming CaCl2 with optimal temperature range of ca. 500 °C–650 °C. However, at higher temperatures (than ca. 700 °C) the retained Cl is released back which is then available for the chlorination of Cd and other HMs again [68,76,77]. Interaction with CaCO3 has also been described in the literature [43] with optimal temperature range of 600–850 °C (Reaction (15)):
C a C O 3 + 2 H C l C a C l 2 + C O 2 + H 2 O
This is consistent with observations of Tang et al. [78] adding CaO or CaCO3 to MSW and evaluating Cd retention in bottom ash. At 700 °C–800 °C, some (moderate) capture of Cd in bottom ash was observed, whereas at 900 °C the release of Cd from bottom ash was even enhanced, due to experimental temperature exceeding the optimal temperature range of 500 °C–650 °C (releasing retained Cl back).
The effect of temperature evaluated by Liu et al. [43] led to a similar conclusion. Sewage sludge conditioned with FeCl3/CaO (due to dewatering) was incinerated at 600 °C–700 °C released slightly less Cd than without conditioner. However, at 800 °C–900 °C, sludge with FeCl3/CaO released more Cd [43]—the latter case corresponds with the temperature exceeding the optimal retention range.
According to Li et al. [52], the retention ability of CaCO3 can be improved by K2CO3 or Al2(SO4)3 modification—the defects created during limestone modification enhanced the retention ability.
Results from recent studies indicate that the presence of PVC during the combustion/incineration of coal/waste with Cd increases its volatility and potential fractions released in emissions. Calcareous additives could suppress this unfavorable effect; but if added directly into the furnace, the temperature is usually too high for efficient retention of chlorination agents [43,52,68,78,79]. However, if limestone is applied within a semi-dry spray tower, good Cd retention results were achieved, which was documented by Chen et al. [57]. Synthetic solid waste (sawdust, PP and Cd-nitrate) was incinerated at 700 °C–800 °C and semi-dry spray tower operate at the temperature of 150 °C–170 °C. If no PVC was added in the feed waste, the best (10%) Cd retention results were received by water injection (i.e., no adsorbent). When PVC was added to synthetic waste, absolutely the best overall Cd retention was observed if limestone was used in semi-dry spray tower (nearly 40% Cd retention), followed by water (21%), kaolinite (13%) and Al2O3 (10%). If NaCl was added to the feed waste, limestone injection exhibited somewhat lower Cd retention (28% Cd retention), which was the best result of all adsorbents injected in the experiment with NaCl.
Another promising adsorbent for Cd retention (if combusted/incinerated with PVC) is NaHCO3 [80]. During rotary kiln incineration (850 °C) of limestone (200 kg) doped with 2 kg PVC (with 0.5% Cd2+), overall Cd retention achieved by post-incineration injection of NaHCO3 was 96.3%.
In summary, emissions of Cd (or potentially of other HMs) could be reduced by controlling Cl content in coal but namely in co-combusted wastes. Other promising approach is to apply adsorbent capturing chlorination gases, e.g., calcareous additives (CaO, CaCO3 etc.). In FBC, such additives are already widely applied. However, they are typically added directly to high-temperature zone where the combustion temperature is ca. 850 °C which is too high to retain HCl. Therefore, from the perspective of the efficient retention of Cd (or possibly also other HMs), low-temperature application is expected to provide better retention results.

6. Effect of Moisture

In addition to different moisture levels in individual coals, even more significant effect can be expected if coal is co-combusted with wastes, from which some typically contain higher moisture levels, such as sewage sludge, agricultural or forest residues, etc. [81,82].
The moisture effect should be also taken into account in the case of oxy-fuel combustion [83] or if coal-water slurry is combusted [84,85]. Different drying methods applied to MSW could also have some effect [86].

6.1. Effect of Moisture (No Extra Cl Added)

No noticeable effect of moisture levels on Cd distribution was concluded by Durlak et al. [87] or Morf et al. [88]. Other studies document increasing Cd emissions with increasing moisture levels [89,90,91].
During pyrolysis of real MSW, release of Cd gradually increased with increasing moisture levels from 0% via 30% up to 65% [90]. The overall enhancement of Cd release corresponding to increase of moisture level up to 65% was ca. 10% (from 60% to 70%) and it was attributed to prolonged pyrolysis time thereby enhancing the volatility of Cd (and some other HMs) [90].
During incineration (at 950 °C) of MSW (paper, flour, cotton, sawdust, polystyrene, Al2O3, SiO2 and Cd-acetate) [89,91], original Cd retention rate in bottom ash (ca. 18%) moderately decreased with increasing moisture levels (in the former study [89] moisture level increased up to 62% H2O and in the latter [91] up to 39.4% H2O). Prevailing physical aspect of the moisture effect on Cd volatility was concluded in these two studies as well [89,91] and prolonged evaporation and longer combustion time were evaluated as dominant reasons of increased Cd volatility. In these two studies [89,91], the incineration time was set as 6 min; therefore, any prolongation might result in increased Cd volatility. In contrast, if MSW was incinerated by Morf et al. [88], the incineration time of 2 h was probably long enough to suppress the effect of some further little prolongation due to moisture evaporation. Equilibrium calculations by Durlak et al. [87] evaluated thermodynamic effect of chemical reactions; kinetic effects are not usually included in such studies [92].

6.2. Effect of Moisture in the Presence of Cl

Cd is one of the most volatile elements, which results in typically low remaining fractions in bottom ash after combustion. If extra Cl is added, Cd volatility further increases leaving behind only a few % of Cd in bottom ash (which somewhat hinders evaluation of the moisture effect).
During synthetic MSW incineration at 950 °C without extra Cl added [91], bottom ash retained ca. 18% of Cd while if 1% of Cl was added, there was practically no retention (<ca. 1–2%)—these data are for 0% moisture. If % H2O increased up to 39.4%, the retention of Cd in bottom ash slightly increased but still remained quite low (ca. < 5%). This effect is in line with thermodynamic equilibrium calculations [91] showing two dominant competitive species: CdCl2 (g) and CdO (s). As the fraction of Cd (g) is within only a few %, the distribution of Cd between CdCl2 (g) and CdO (s) governs overall Cd volatility and its release from bottom ash. Specifically, at 950 °C [91], increase of moisture from 10 to 35% lead to decrease of CdCl2 (g) fraction by ca. 20% (and this Cd fraction is almost totally converted to CdO(s) because increase of Cd (g) makes only a few %). And the conversion of gaseous CdCl2 to condensed CdO decreases Cd volatility. Overall, if Cl is present in sufficient amount in the feed fuel, the volatility enhanced by Cl prevails on the retention effect brought about by moisture (there are two opposite effects of Cl and moisture).

7. Effect of Atmosphere

7.1. Effect of O2 Level

The air/coal ratio is an important factor controlling Cd volatility and emissions [68,92]. According to thermodynamic equilibrium calculations simulating pulverized coal combustion, all Cd remained in the solid phase at 900 K if air/coal ratio was at least 1.1; decrease of the air/coal ratio to 1.0 resulted in all Cd to be in the vapor phase (even at 600 K). In the case of FBC, an increase of the air/coal ratio from 1.0 to 1.1 changed Cd distribution significantly toward its enhanced fractions in the solid phase. Nevertheless, these calculations are based on reaching the chemical equilibrium.
As for the experimental results, it is noteworthy that (at least) 3 different conclusions about effect of % O2 on Cd volatility can be found in the literature [93,94,95]:
(i) Decrease of Cd volatility by increased % O2 (which is consistent with thermodynamic calculations [92]) was concluded for incineration of Al2O3 impregnated with CdCl2 at FBC reactor at 850 °C for 1 h [93]. The observation was explained via the Reaction (16) [93]:
2 M C l + A l 2 O 3 + 2 O 2 2 M O 2 · A l 2 O 3 + 2 C l 2
(where increasing % O2 shifts the equilibrium to the right in favor of the formation of non-volatile 2 M O 2 · A l 2 O 3 ) [93].
As the dwell time in the high-temperature zone was long enough (1 h), the kinetic limitations were thereby diminished and the experimental reasons are in line with thermodynamic equilibrium calculations.
Thermodynamic equilibrium calculations conducted by Yan et al. [62] concluded different effect of oxidizing/reducing conditions for low-ash coal and high-ash coal. In the case of high-ash coal, volatilization of Cd in different atmospheres occurs in similar temperature regions (oxidizing atmosphere at 550–800 K; reducing at 600–800 K) while in the case of low-ash coal, volatilization in oxidizing atmosphere occurs at higher temperatures (800–1000 K) than in the case of reducing conditions (600–700 K). In this case, reducing conditions facilitate Cd volatilization.
(ii) Increasing Cd volatility with increasing % O2 in the inlet atmosphere was observed during co-combustion of lignite and waste activated sludge for 30 min at 1000 °C in horizontal tube furnace [94]. Vaporization percentage of Cd increased from ca. 35% (10% O2) up to ca. 50% (30% O2) and it was explained through enhanced intensity of combustion by higher % O2 (increasing local temperature to some extent).
(iii) No noticeable effect of % O2 on Cd volatilization was concluded during the combustion of two coals at 21% O2 (in N2) and 6.35% O2 (in N2) atmospheres at 800 and 900 °C [95]. The crucible containing the coal was lifted up and down inside the vertical furnace and no matter which temperature was used, % O2 did not affect Cd volatility in these two studied coals. It should be mentioned that the weight loss of Cd (from the coal) was quite low: ca. 35% was volatilized at both temperatures and O2 levels from the first coal; in the case of the second coal, at 800 °C ca. 25% (for both O2 levels) and at 900 °C ca. 50% of Cd was released. As Cd is a typical high-volatility element, the aforementioned volatilized fractions are surprisingly low, which might correspond with the very character of these combustion experiments, as they were conducted with the aim to simulate the initial stage of flame quenching (i.e., during the time needed only for the release of volatile matter, its ignition and burnout). Therefore, some Cd still remained in the rest of the sample and it could be hypothesized that O2 level is more important for the interaction with the rest of the sample (than for the volatiles burnout).

7.2. Effect of “the Rest” of the Atmosphere

Not only % O2, but also the composition of “the rest” of the combustion/incineration atmosphere can play an important role [78,96], which is studied namely due to oxy-fuel combustion technology where CO2 can be captured due to recycling of the flue gas and oxygen is used instead of air (thereby increasing CO2 content up to 95%) [96]. Thus, instead of N2/O2 atmosphere, CO2/O2 is used in oxy-fuel combustion.
Thermodynamic equilibrium calculations [97] concluded that distribution of Cd (and other HMs) in oxy-fuel combustion was nearly the same as in the case of air combustion. However, real incineration experiments with synthetic MSW [78] (flour, paper, sawdust, HDPE, texture, rubber, leather and Cd-acetate) in tubular furnace revealed certain differences in Cd volatility in CO2/O2 and N2/O2 atmospheres. In both atmospheres, the Cd residual rates at 1000 °C were almost zero; at 900 °C they were nearly the same—ca. 27–28%. However, at 700 °C and 800 °C, Cd residual rates were higher in CO2/O2 (oxy-fuel atmosphere) than in N2/O2 (air): ca. 68% for CO2/O2 and 52% for N2/O2 at 700 °C and 25% (CO2/O2) and 22% (N2/O2) for 800 °C. The suppressed Cd volatility in CO2/O2 was more pronounced at lower temperatures, which was explained through higher specific heat capacity of CO2 than N2 resulting in delaying the ignition of the MSW and lowering the combustion temperature (in the case of CO2/O2 atmosphere) thereby decreasing the emissions of Cd. As the incineration time was the same for both atmospheres, these results are not contradictory to thermodynamic equilibrium calculations [97] where kinetic limitations are not usually taken into account. Lower volatilization rate of HMs as a result of a lower particle combustion temperature in oxy-fuel combustion (hence affecting the elemental enrichment in ash) was concluded also by Roy and Bhattacharya [98], Font et al. [99], Oboirien et al. [100] or Suriyawong et al. [101].

8. Effect of S

Thermodynamic equilibrium calculations (under oxidizing conditions) concluded the occurrence of CdSO4 up to ca. 730 °C [102]–750 °C [103,104]. Some authors evaluated the maximum temperature of the predominant CdSO4 occurrence to be ca. 600 [37] or 630 °C [48] (these minor differences probably originate from different interacting components, atmosphere composition and concentrations of Cd and SO42−). In any case, even during FBC in power stations (where the combustion temperature is quite low—ca. 850 °C), CdSO4 (s) is not stable. If the temperature exceeds 600–700 °C (up to ca. 1000 °C), there is a coexistence of multiple phases whose composition is dependent on levels of interacting components. If the presence of chlorides, silica and alumina (along with S) is taken into account, there are namely CdCl2 (s), CdAl2O4 (s), CdSiO3 (s) and CdO (s). Then, the highest temperature (t > ca. 1000–1100 °C) is characterized by prevalent occurrence of Cd (g) as the aforementioned species are not stable in this high-temperature region [102,105,106].
At lower temperatures (t < ca. 600–700 °C), Cd is not the only element that binds sulfates—there are also Na2SO4 and CaSO4 (as Na and Ca are common elements in coal, wastes or desulfurization additives) or Cr2(SO4)3, HgSO4, PbSO4 etc. [48]. Individual metals bind S according to their affinities; according to Yao and Naruse [48], if concentration of S is low, alkali metals are preferred to form condensed phases with it and with increasing S concentration, Cd followed by Pb, Hg and Cr create their sulfates as well.
Binding energies calculated for interaction between Cd and S2−, Cl, sulfate and phosphate document quite high values for sulfates and phosphates (in comparison with sulfides and chlorides); it means that is case of co-existence of all these species, the formation of sulfides or chlorides is preferred [69]. This is consistent with moderate increase of % Cd in bottom ash during laboratory combustion (at 850 °C) of synthetic MSW with Cd-acetate and S or Na2S added [107].
In any case, theoretic equilibrium calculations indicate that oxidizing conditions and higher temperatures are not favorable for efficient retention of Cd by S, which is consistent (e.g.,) with experimental results presented by Luan et al. [69] where S added to wastewater sludge combusted at 850 °C did not changed noticeably the volatilization rate of Cd (which was ca. 40–41% no matter if S was added).
Decrease of Cd retention by SO2 injection during co-combustion of coal, biomass and waste secondary fuels (in FBC reactor) was reported in the literature as well [37,57,107]. Thermal instability of CdSO4 in combination with the combustion temperature of 850 °C (in laboratory combustor used for synthetic MSW) might be the probable reason why the addition of Na2SO4 increased % Cd in flue gas [101].
Nevertheless, there are also studies documenting that adding Na2SO4 might favour the retention of Cd [46,57]. Chen et al. [46] combusted synthetic solid waste (plastics, sawdust and Cd-nitrate) at 700 °C in reactor simulating FBC along with 3 different sorbents—kaolinite, bauxite and Al2O3 (added into the combustion chamber). Retention rate of Cd on kaolinite, bauxite and Al2O3 was ca. 8%, 13% and 8%, respectively. If Na2SO4 was added into the combustion chamber (under the same conditions), Cd retention efficiency increased up to 25%, 21% and 34%. It means that the most efficient capture of Cd was achieved by adding Na2SO4 and Al2O3 to synthetic solid waste (directly at combustion chamber with 700 °C). Without Na2SO4, the best retention was achieved by bauxite (ca. 13%), which is still quite low in comparison with combined effect of Na2SO4–Al2O3.
Chen et al. [57] published later the results of other similar experiment. Synthetic solid waste (plastics, sawdust and Cd-nitrate) was combusted at 700 °C in the reactor simulating fluidized-bed incineration. In this case, a semi-dry spray tower (150–170 °C) was installed for injection of water with 4% of kaolinite, Al2O3, limestone or without any adsorbent. The best overall retention of Cd was observed if no adsorbent (i.e., only H2O injection in semi-dry spray tower) was used—ca. 10%. When Na2SO4 was added to the feed waste, overall Cd retention increased up to 50% if combined with limestone injection. If no adsorbent was used, Na2SO4 itself increased the Cd retention from ca. 10% to ca. 13%. If limestone adsorbent was used without addition of Na2SO4 to the feed waste, the retention was even worse (<5%). Therefore, there is a synergic effect of Na2SO4 and limestone that efficiently increase Cd retention.
Yu et al. [93] observed significant increase of Cd vaporization when SO2 was added into N2 atmosphere in FBC reactor (at 850 °C) during combustion of CdCl2-impregnated porous Al2O3 matrix. After 30-min incineration time ca. 30% Cd was vaporized in pure N2; if 2000 ppm SO2 and 5000 ppm SO2 was present, vaporization percentage of Cd increased up to 90% and 95%. It is interesting that the effect on Pb was much lower (vaporization did not exceed 20%) and Zn and Cu did not show any significant evaporation.
It is known that sulfate phase may be formed at low temperature in the presence of S; at higher temperatures it does not play a significant role as CdSO4 is decomposed at t > 850 °C [93,108].

9. Effect of P

Binding energy for interaction between Cd and phosphate is the highest from the energies related to Cl, S2− and sulfate. Natural levels of P in most coals are low (200–250 ppm) [22]. Nevertheless, the major advantage of phosphates is their high thermal stability (unlike e.g., sulfates).
If P content in wastewater sludge was increased from 3.2% to 7% P (in the form of P2O5), Cd volatilization at 850 °C decreased from 40% to ca. 17% [69]. Promising results were presented also for sediment combustion (at 600 °C and 850 °C) previously treated with phosphoric acid (4.50 PO43−). Cd vaporization % almost does not change with temperature increase from 600 to 850 °C (15 and 17% vaporized) [80].

10. Conclusions

Concentrations of Cd in coal are quite low (0.033–0.64 ppm) [23]; however, (e.g.,) in MSW or municipal wastewater sludge it can be significantly higher: 0–90 and ca. 100 ppm, respectively [23]. In contrast, desulfurization additives as limestone (ca. 0.01 ppm Cd) or urea (ca. 0.03 ppm Cd) might exhibit diluting effect on Cd concentrations [25]. In coal, Cd is associated with mono/disulfides with some minor occurrence in organic matter or (alumino)silicates. Due to high Cd volatility (with exception of typically minor (alumino)silicate-bound fraction), most Cd is easily volatilized and exhibit potential risk of being emitted to surrounding atmosphere either in gaseous form or in sub-micron particles. Therefore, particular attention is paid to mitigating these emissions.
Different approaches for decreasing emissions could be used: pre-combustion treatment, mid-combustion or post-combustion control [52].
Within the pre-combustion treatment, increase of Cd retention was achieved by H3PO4 treatment [69].
For their use in high-temperature (i.e., in-furnace) regime, kaolinite [54] and bauxite [50] exhibited promising results at 1100–1200 °C. Moreover, Cd retained within high-temperature capture (typically by melted or at least agglomerated particles) should be more resistant to leaching than Cd physically sorbed/condensed during low-temperature regime. For the commonly used desulfurization adsorbents (CaO, limestone etc.) exhibiting efficient retention of S and other elements (e.g., As) [8], even FBC temperature might be too high to provide efficient retention of chlorination agents that otherwise enhance Cd (and some other HMs) volatility. Optimal temperature range in this case is 500–700 °C [43,68,75,76]. Lower temperatures typically also favor Cd retention by S [37,48,102].
Post-combustion treatment of flue gas (using adsorbents in low-temperature regime) has also been tested. Promising results has been achieved using semi-dry spray tower working at 150–170 °C [57] spraying suspension of 5% adsorbent (in water). The adsorbents were efficient for most HMs but in the case of Cd the best results have been observed if water without adsorbents was used.
Detailed information related to thermodynamic distribution of individual Cd (or other HMs) species in dependence on temperature, atmosphere, interacting-species concentrations etc. can be found in the literature. These calculations are often in line with laboratory experiments, namely if combustion/incineration/adsorption dwell time is long enough. Nevertheless, as typical coal combustion in power stations provides (at maximum) a few minutes in high-temperature zone, future trend (in mitigating HMs emissions) could be directed toward the evaluation of time factor, i.e., the kinetic aspect of the HMs volatilization/retention.
In any case, using pre-combustion treatment or in-furnace additives/adsorbents changes the composition and other characteristics of combustion (incineration) ashes, which may significantly affect their further technological utilization or environmental impact (e.g., leaching). Similarly, using low-temperature adsorbents may either affect the fly ash quality or even create new wastes. All these alternatives should be taken into account as studies evaluating these aspects are currently quite scarce. The effect on fouling/slagging could be evaluated as well. Mitigating emissions of HMs is not the only challenge related to the coal combustion; new technologies are developed to decrease emissions of CO2, NOx and other pollutants and distribution of HMs under these conditions should be evaluated as well.

Author Contributions

Conceptualization, literature search and evaluation, writing and editing, L.B.; funding acquisition, project administration and consultations during manuscript preparation, H.R. and M.K.; consultations on industrial coal combustion, B.Č. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education, Youth and Sport of the Czech Republic by the research projects: CZ.1.05/2.1.00/19.0389: Research Infrastructure Development of the CENET and SP2020/22 “Innovative methods for monitoring particulate matter from combustion processes”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Parzentny, H.R.; Róg, L. The Role of Mineral Matter in Concentrating Uranium and Thorium in Coal and Combustion Residues from Power Plants in Poland. Minerals 2019, 9, 312. [Google Scholar] [CrossRef] [Green Version]
  2. Wagner, N.; Matiane, A. Rare earth elements in select Main Karoo Basin (South Africa) coal and coal ash samples. Int. J. Coal Geol. 2018, 196, 82–92. [Google Scholar] [CrossRef]
  3. Hower, J.C.; Granite, E.J.; Mayfield, D.B.; Lewis, A.S.; Finkelman, R.B. Notes on Contributions to the Science of Rare Earth Element Enrichment in Coal and Coal Combustion Byproducts. Minerals 2016, 6, 32. [Google Scholar] [CrossRef] [Green Version]
  4. Bartoňová, L.; Serenčíšová, J.; Čech, B. Yttrium partitioning and associations in coal-combustion ashes prior to and after their leaching in HCl. Fuel Process. Technol. 2018, 173, 205–215. [Google Scholar] [CrossRef]
  5. Klika, Z.; Ambružová, L.; Sýkorová, I.; Seidlerová, J.; Kolomazník, I. Critical evaluation of sequential extraction and sink-float methods used for the determination of Ga and Ge affinity in lignite. Fuel 2009, 88, 1834–1841. [Google Scholar] [CrossRef]
  6. Raclavská, H.; Matýsek, D.; Raclavský, K.; Juchelková, D. Geochemistry of fly ash from desulphurisation process performed by sodium bicarbonate. Fuel Process. Technol. 2010, 91, 150–157. [Google Scholar] [CrossRef]
  7. Scala, F.; Chirone, R.; Meloni, P.; Carcangiu, G.; Manca, M.; Mulas, G.; Mulas, A. Fluidized bed desulfurization using lime obtained after slow calcination of limestone particles. Fuel 2013, 114, 99–105. [Google Scholar] [CrossRef]
  8. Bartoňová, L.; Klika, Z. Effect of CaO on retention of S, Cl, Br, As, Mn, V, Cr, Ni, Cu, Zn, W and Pb in bottom ashes from fluidized-bed coal combustion power station. J. Environ. Sci. 2014, 26, 1429–1436. [Google Scholar] [CrossRef]
  9. Scala, F.; Cimino, S. Elemental mercury capture and oxidation by a regenerable manganese-based sorbent: The effect of gas composition. Chem. Eng. J. 2015, 278, 134–139. [Google Scholar] [CrossRef]
  10. Bartoňová, L. Unburned carbon from coal combustion ash: An overview. Fuel Process. Technol. 2015, 134, 136–158. [Google Scholar] [CrossRef]
  11. Hower, J.C.; Groppo, J.G.; Graham, U.M.; Ward, C.R.; Kostova, I.J.; Maroto-Valer, M.M.; Dai, S. Coal-derived unburned carbons in fly ash: A review. Int. J. Coal Geol. 2017, 179, 11–27. [Google Scholar] [CrossRef]
  12. Wagner, N.; Hlatshwayo, B. The occurrence of potentially hazardous trace elements in five Highveld coals, South Africa. Int. J. Coal Geol. 2005, 63, 228–246. [Google Scholar] [CrossRef]
  13. Scala, F.; Chirone, R.; Lancia, A. Elemental mercury vapor capture by powdered activated carbon in a fluidized bed reactor. Fuel 2011, 90, 2077–2082. [Google Scholar] [CrossRef]
  14. Hower, J.C.; Robl, T.; Anderson, C.; Thomas, G.; Sakulpitakphon, T.; Mardon, S.; Clark, W. Characteristics of coal combustion products (CCP’s) from Kentucky power plants, with emphasis on mercury content. Fuel 2005, 84, 1338–1350. [Google Scholar] [CrossRef]
  15. Senior, C.L.; Johnson, S.A. Impact of Carbon-in-Ash on Mercury Removal across Particulate Control Devices in Coal-Fired Power Plants. Energy Fuels 2005, 19, 859–863. [Google Scholar] [CrossRef]
  16. Antón, M.A.L.; Diaz-Somoano, M.; Fierro, J.; Martinez-Tarazona, M.R. Retention of arsenic and selenium compounds present in coal combustion and gasification flue gases using activated carbons. Fuel Process. Technol. 2007, 88, 799–805. [Google Scholar] [CrossRef] [Green Version]
  17. Seames, W.S.; Wendt, J.O. Partitioning of arsenic, selenium, and cadmium during the combustion of Pittsburgh and Illinois #6 coals in a self-sustained combustor. Fuel Process. Technol. 2000, 63, 179–196. [Google Scholar] [CrossRef]
  18. Bartoňová, L.; Raclavská, H.; Čech, B.; Kucbel, M. Behavior of Pb During Coal Combustion: An Overview. Sustainability 2019, 11, 6061. [Google Scholar] [CrossRef] [Green Version]
  19. Tun, M.M.; Juchelková, D.; Raclavská, H.; Sassmanová, V. Utilization of Biodegradable Wastes as a Clean Energy Source in the Developing Countries: A Case Study in Myanmar. Energies 2018, 11, 3183. [Google Scholar] [CrossRef] [Green Version]
  20. Hoornweg, D.; Bhada-Tata, P. What a Waste: A Global Review of Solid Waste Management; Urban Development Series Knowledge Papers; The World Bank Group: Washington, DC, USA, 2012; pp. 1–98. [Google Scholar]
  21. Tun, M.M.; Juchelková, D. Assessment of solid waste generation and greenhouse gas emission potential in Yangon city, Myanmar. J. Mater. Cycles Waste Manag. 2018, 20, 1397–1408. [Google Scholar] [CrossRef]
  22. Ketris, M.; Yudovich, Y. Estimations of Clarkes for Carbonaceous biolithes: World averages for trace element contents in black shales and coals. Int. J. Coal Geol. 2009, 78, 135–148. [Google Scholar] [CrossRef]
  23. Biswas, P.; Wu, C.Y. Control of toxic metal emissions from combustors using sorbents: A review. J. Air Waste Manag. Assoc. 1998, 48, 113–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Nowak, B.; Aschenbrenner, P.; Winter, F. Heavy metal removal from sewage sludge ash and municipal solid waste fly ash—A comparison. Fuel Process. Technol. 2013, 105, 195–201. [Google Scholar] [CrossRef]
  25. Fu, B.; Liu, G.; Sun, M.; Hower, J.C.; Mian, M.; Wu, D.; Wang, R.; Hu, G. Emission and transformation behavior of minerals and hazardous trace elements (HTEs) during coal combustion in a circulating fluidized bed boiler. Environ. Pollut. 2018, 242, 1950–1960. [Google Scholar] [CrossRef]
  26. Fu, B.; Liu, G.; Sun, M.; Hower, J.C.; Hu, G.; Wu, D. A comparative study on the mineralogy, chemical speciation, and combustion behavior of toxic elements of coal beneficiation products. Fuel 2018, 228, 297–308. [Google Scholar] [CrossRef]
  27. Wang, Y.; Tang, Y.; Liu, S.; Wang, Y.; Finkelman, R.B.; Wang, B.; Guo, X. Behavior of trace elements and mineral transformations in the super-high organic sulfur Ganhe coal during gasification. Fuel Process. Technol. 2018, 177, 140–151. [Google Scholar] [CrossRef]
  28. Finkelman, R.B.; Palmer, C.A.; Wang, P. Quantification of the modes of occurrence occurrence of 42 elements in coal. Int. J. Coal Geol. 2018, 185, 138–160. [Google Scholar] [CrossRef]
  29. Wagner, N.; Tlotleng, M. Distribution of selected trace elements in density fractionated Waterberg coals from South Africa. Int. J. Coal Geol. 2012, 94, 225–237. [Google Scholar] [CrossRef]
  30. Parzentny, H.R.; Róg, L. Distribution of Some Ecotoxic Elements in Fuel and Solid Combustion Residues in Poland. Energies 2020, 13, 1131. [Google Scholar] [CrossRef] [Green Version]
  31. Yudovich, Y.; Ketris, M. Chlorine in coal: A review. Int. J. Coal Geol. 2006, 67, 127–144. [Google Scholar] [CrossRef]
  32. Vassilev, S. Contents, modes of occurrence and origin of chlorine and bromine in coal. Fuel 2000, 79, 903–921. [Google Scholar] [CrossRef]
  33. Ohenoja, K.; Pesonen, J.; Yliniemi, J.; Illikainen, M. Utilization of Fly Ashes from Fluidized Bed Combustion: A Review. Sustainability 2020, 12, 2988. [Google Scholar] [CrossRef] [Green Version]
  34. Vassilev, S.V.; Vassileva, C.G.; Vassilev, V.S. Advantages and disadvantages of composition and properties of biomass in comparison with coal: An overview. Fuel 2015, 158, 330–350. [Google Scholar] [CrossRef]
  35. Růžičková, J.; Raclavská, H.; Kucbel, M.; Grobelak, A.; Šafář, M.; Raclavský, K.; Švédová, B.; Juchelková, D.; Moustakas, K. The potential environmental risks of the utilization of composts from household food waste. Environ. Sci. Pollut. Res. 2020, 1–17. [Google Scholar] [CrossRef]
  36. Corsaro, A.; Raclavská, H.; Hlavsová, A.; Frydrych, J.; Juchelková, D. Perennial grasses as prospective energy sources. Energy Sources Part A Recover. Util. Environ. Eff. 2016, 38, 1206–1211. [Google Scholar] [CrossRef]
  37. Diaz-Somoano, M.; Unterberger, S.; Hein, K. Prediction of trace element volatility during co-combustion processes. Fuel 2006, 85, 1087–1093. [Google Scholar] [CrossRef]
  38. Youcai, Z.; Stucki, S.; Ludwig, C.; Wochele, J. Impact of moisture on volatility of heavy metals in municipal solid waste incinerated in a laboratory scale simulated incinerator. Waste Manag. 2004, 24, 581–587. [Google Scholar] [CrossRef]
  39. Kovacs, H.; Szemmelveisz, K.; Koós, T. Theoretical and experimental metals flow calculations during biomass combustion. Fuel 2016, 185, 524–531. [Google Scholar] [CrossRef]
  40. Perry, D.L. Handbook of Inorganic Compounds, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2011; p. 581. ISBN 9781439814611. [Google Scholar]
  41. Chan, C.; Jia, C.Q.; Graydon, J.W.; Kirk, D.W. The behaviour of selected heavy metals in MSW incineration electrostatic precipitator ash during roasting with chlorination agents. J. Hazard. Mater. 1996, 50, 1–13. [Google Scholar] [CrossRef]
  42. Yu, J.; Qiao, Y.; Jin, L.; Ma, C.; Paterson, N.; Sun, L. Removal of toxic and alkali/alkaline earth metals during co-thermal treatment of two types of MSWI fly ashes in China. Waste Manag. 2015, 46, 287–297. [Google Scholar] [CrossRef]
  43. Liu, J.; Zeng, J.; Sun, S.; Huang, S.; Kuo, J.; Chen, N. Combined effects of FeCl3 and CaO conditioning on SO2, HCl and heavy metals emissions during the DDSS incineration. Chem. Eng. J. 2016, 299, 449–458. [Google Scholar] [CrossRef]
  44. Linak, W.P.; Wendt, J.O. Toxic metal emissions from incineration: Mechanisms and control. Prog. Energy Combust. Sci. 1993, 19, 145–185. [Google Scholar] [CrossRef]
  45. Scotto, M.V.; Uberoi, M.; Peterson, T.W.; Shadman, F.; Wendt, J.O. Metal capture by sorbents in combustion processes. Fuel Process. Technol. 1994, 39, 357–372. [Google Scholar] [CrossRef]
  46. Chen, J.-C.; Wey, M.-Y.; Lin, Y.-C. The adsorption of heavy metals by different sorbents under various incineration conditions. Chemosphere 1998, 37, 2617–2625. [Google Scholar] [CrossRef]
  47. Ho, T.C.; Chuang, T.C.; Chelluri, S.; Lee, Y.; Hopper, J.R. Simultaneous capture of metal, sulfur and chlorine by sorbents during fluidized bed incineration. Waste Manag. 2001, 21, 435–441. [Google Scholar] [CrossRef]
  48. Yao, H.; Naruse, I. Using sorbents to control heavy metals and particulate matter emission during solid fuel combustion. Particuology 2009, 7, 477–482. [Google Scholar] [CrossRef]
  49. Yao, H.; Mkilaha, I.S.; Naruse, I. Screening of sorbents and capture of lead and cadmium compounds during sewage sludge combustion. Fuel 2004, 83, 1001–1007. [Google Scholar] [CrossRef]
  50. Cheng, J.F.; Zeng, H.C.; Zhang, Z.H.; Xu, M.H. The effects of solid absorbents on the emission of trace elements, SO2, and NOx during coal combustion. Int. J. Energy Res. 2001, 25, 1043–1052. [Google Scholar] [CrossRef]
  51. Chen, J.; Sun, Y.; Shao, N.; Zhang, Z. Environmental mitigation of sludge combustion via two opposite modifying strategies: Kinetics and stabilization effect. Fuel 2018, 227, 346–354. [Google Scholar] [CrossRef]
  52. Li, S.; Guo, S.; Huang, X.; Huang, T.; Niazi, N.K.; Muhammad, F.; Xu, G.; Zhao, Z.; Yu, L.; Yan, Y.; et al. Research on characteristics of heavy metals (As, Cd, Zn) in coal from Southwest China and prevention method by using modified calcium-based materials. Fuel 2016, 186, 714–725. [Google Scholar] [CrossRef]
  53. Kuo, J.-H.; Lin, C.-L.; Wey, M.-Y. Effect of particle agglomeration on heavy metals adsorption by Al- and Ca-based sorbents during fluidized bed incineration. Fuel Process. Technol. 2011, 92, 2089–2098. [Google Scholar] [CrossRef]
  54. Wendt, J.O.L.; Lee, S.J. High-temperature sorbents for Hg, Cd, Pb, and other trace metals: Mechanisms and applications. Fuel 2010, 89, 894–903. [Google Scholar] [CrossRef]
  55. Peng, T.-H.; Lin, C.-L. Influence of various chlorine additives on the partitioning of heavy metals during low-temperature two-stage fluidized bed incineration. J. Environ. Manag. 2014, 146, 362–368. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, J.-C.; Wey, M.-Y.; Yan, M.-H. Theoretical and Experimental Study of Metal Capture during Incineration Process. J. Environ. Eng. 1997, 123, 1100–1106. [Google Scholar] [CrossRef]
  57. Chen, J.-C.; Wey, M.-Y.; Ou, W.-Y. Capture of heavy metals by sorbents in incineration flue gas. Sci. Total Environ. 1999, 228, 67–77. [Google Scholar] [CrossRef]
  58. Chiang, B.-C.; Wey, M.-Y.; Yeh, C.-L. Control of acid gases using a fluidized bed adsorber. J. Hazard. Mater. 2003, 101, 259–272. [Google Scholar] [CrossRef]
  59. Wey, M.-Y.; Chen, K.-H.; Liu, K.-Y. The effect of ash and filter media characteristics on particle filtration efficiency in fluidized bed. J. Hazard. Mater. 2005, 121, 175–181. [Google Scholar] [CrossRef]
  60. Folgueras, M.B.; Folgueras-Díaz, M.; Xiberta, A.J.; Alonso, M. Effect of Inorganic Matter on Trace Element Behavior during Combustion of Coal−Sewage Sludge Blends. Energy Fuels 2007, 21, 744–755. [Google Scholar] [CrossRef]
  61. Danihelka, P.; Volna, Z.; Jones, J.; Williams, A. Emission of trace toxic metals during pulverized fuel combustion of Czech coals. Int. J. Energy Res. 2003, 27, 1181–1203. [Google Scholar] [CrossRef]
  62. Yan, R.; Gauthier, D.; Flamant, G. Volatility and chemistry of trace elements in a coal combustor. Fuel 2001, 80, 2217–2226. [Google Scholar] [CrossRef]
  63. Raclavská, H.; Corsaro, A.; Hartmann-Koval, S.; Juchelková, D. Enrichment and distribution of 24 elements within the sub-sieve particle size distribution ranges of fly ash from wastes incinerator plants. J. Environ. Manag. 2017, 203, 1169–1177. [Google Scholar] [CrossRef] [PubMed]
  64. Chiang, K.-Y.; Wang, K.-S.; Lin, F.-L.; Chu, W.-T. Chloride effects on the speciation and partitioning of heavy metal during the municipal solid waste incineration process. Sci. Total Environ. 1997, 203, 129–140. [Google Scholar] [CrossRef]
  65. Huang, Y.J.; Jin, B.S.; Zhong, Z.P.; Xiao, R.; Tang, Z.Y.; Ren, H.F. Emission features of several trace elements in pulverized coal boiler. In Proceedings of the International Conference on Energy and the Environment, Shanghai, China, 11–13 December 2003; Volume 1–2, pp. 562–568, ISBN 7-5323-7335-5. [Google Scholar]
  66. Wu, H.; Glarborg, P.; Frandsen, F.J.; Dam-Johansen, K.; Jensen, P.A.; Sander, B. Trace elements in co-combustion of solid recovered fuel and coal. Fuel Process. Technol. 2013, 105, 212–221. [Google Scholar] [CrossRef] [Green Version]
  67. Qi, X.; Song, G.; Yang, S.; Yang, Z.; Lyu, Q. Migration and transformation of sodium and chlorine in high-sodium high-chlorine Xinjiang lignite during circulating fluidized bed combustion. J. Energy Inst. 2019, 92, 673–681. [Google Scholar] [CrossRef]
  68. Wang, X.; Huang, Y.; Liu, C.; Zhang, S.; Wang, Y.; Piao, G. Dynamic volatilization behavior of Pb and Cd during fixed bed waste incineration: Effect of chlorine and calcium oxide. Fuel 2017, 192, 1–9. [Google Scholar] [CrossRef]
  69. Luan, J.; Li, R.; Zhang, Z.; Li, Y.; Zhao, Y. Influence of chlorine, sulfur and phosphorus on the volatilization behavior of heavy metals during sewage sludge thermal treatment. Waste Manag. Res. 2013, 31, 1012–1018. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, K.-S.; Chiang, K.-Y.; Tsai, C.C.; Sun, C.-J.; Lin, K.-L.; Tsai, C.-C.; Tsai, C.-C. The effects of FeCl3 on the distribution of the heavy metals Cd, Cu, Cr, and Zn in a simulated multimetal incineration system. Environ. Int. 2001, 26, 257–263. [Google Scholar] [CrossRef]
  71. Nowak, B.; Rocha, S.F.; Aschenbrenner, P.; Rechberger, H.; Winter, F. Heavy metal removal from MSW fly ash by means of chlorination and thermal treatment: Influence of the chloride type. Chem. Eng. J. 2012, 179, 178–185. [Google Scholar] [CrossRef]
  72. Yan, R.; Gauthier, D.; Flamant, G. Partitioning of trace elements in the flue gas from coal combustion. Combust. Flame 2001, 125, 942–954. [Google Scholar] [CrossRef]
  73. Yu, J.; Sun, L.; Ma, C.; Qiao, Y.; Yao, H. Thermal degradation of PVC: A review. Waste Manag. 2016, 48, 300–314. [Google Scholar] [CrossRef]
  74. Saeed, L.; Tohka, A.; Haapala, M.; Zevenhoven, R. Pyrolysis and combustion of PVC, PVC-wood and PVC-coal mixtures in a two-stage fluidized bed process. Fuel Process. Technol. 2004, 85, 1565–1583. [Google Scholar] [CrossRef]
  75. Chyang, C.-S.; Han, Y.-L.; Zhong, Z.-C. Study of HCl Absorption by CaO at High Temperature. Energy Fuels 2009, 23, 3948–3953. [Google Scholar] [CrossRef]
  76. Bie, R.; Li, S.; Yang, L. Reaction mechanism of CaO with HCl in incineration of wastewater in fluidized bed. Chem. Eng. Sci. 2005, 60, 609–616. [Google Scholar] [CrossRef]
  77. Zhu, H.; Jiang, X.; Yan, J.; Chi, Y.; Cen, K. TG-FTIR analysis of PVC thermal degradation and HCl removal. J. Anal. Appl. Pyrolysis 2008, 82, 1–9. [Google Scholar] [CrossRef]
  78. Tang, X.; Chen, F.; Shao, D.; Qin, P. Effects of CaO and CaCO3 on Heavy Metal Capture in Bottom Ash during Municipal Solid Waste Combustion under a CO2/O2 Atmosphere. Energy Fuels 2017, 31, 10998–11006. [Google Scholar] [CrossRef]
  79. Font, O.; Querol, X.; Izquierdo, M.; Alvarez, E.; Moreno, N.; Díez, S.; Álvarez-Rodríguez, R.; Clemente-Jul, C.; Coca, P.; Garcia-Peña, F. Partitioning of elements in a entrained flow IGCC plant: Influence of selected operational conditions. Fuel 2010, 89, 3250–3261. [Google Scholar] [CrossRef]
  80. Rio, S.; Verwilghen, C.; Ramaroson, J.; Ange, N.; Sharrock, P. Heavy metal vaporization and abatement during thermal treatment of modified wastes. J. Hazard. Mater. 2007, 148, 521–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Raclavská, H.; Juchelková, D.; Roubíček, V.; Matýsek, D. Energy utilisation of biowaste—Sunflower-seed hulls for co-firing with coal. Fuel Process. Technol. 2011, 92, 13–20. [Google Scholar] [CrossRef]
  82. Raclavská, H.; Juchelková, D.; Škrobánková, H.; Wiltowski, T.; Campen, A. Conditions for energy generation as an alternative approach to compost utilization. Environ. Technol. 2011, 32, 407–417. [Google Scholar] [CrossRef]
  83. Roy, B.; Choo, W.L.; Bhattacharya, S. Prediction of distribution of trace elements under Oxy-fuel combustion condition using Victorian brown coals. Fuel 2013, 114, 135–142. [Google Scholar] [CrossRef]
  84. Vershinina, K.Y.; Egorov, R.I.; Strizhak, P.A. The ignition parameters of the coal-water slurry droplets at the different methods of injection into the hot oxidant flow. Appl. Therm. Eng. 2016, 107, 10–20. [Google Scholar] [CrossRef]
  85. Glushkov, D.O.; Shabardin, D.P.; Strizhak, P.A.; Vershinina, K.Y. Influence of organic coal-water fuel composition on the characteristics of sustainable droplet ignition. Fuel Process. Technol. 2016, 143, 60–68. [Google Scholar] [CrossRef]
  86. Tun, M.M.; Juchelková, D. Drying methods for municipal solid waste quality improvement in the developed and developing countries: A review. Environ. Eng. Res. 2018, 24, 529–542. [Google Scholar] [CrossRef] [Green Version]
  87. Durlak, S.K.; Biswas, P.; Shi, J. Equilibrium analysis of the affect of temperature, moisture and sodium content on heavy metal emissions from municipal solid waste incinerators. J. Hazard. Mater. 1997, 56, 1–20. [Google Scholar] [CrossRef]
  88. Morf, L.S.; Brunner, P.H.; Spaun, S. Effect of operating conditions and input variations on the partitioning of metals in a municipal solid waste incinerator. Waste Manag. Res. 2000, 18, 4–15. [Google Scholar] [CrossRef]
  89. Meng, A.; Li, Q.; Jia, J.; Zhang, Y. Effect of Moisture on Partitioning of Heavy metals in Incineration of Municipal Solid Waste. Chinese J. Chem. Eng. 2012, 20(5), 1008–1015. [Google Scholar] [CrossRef]
  90. Raclavská, H.; Corsaro, A.; Hlavsová, A.; Juchelková, D.; Zajonc, O. The effect of moisture on the release and enrichment of heavy metals during pyrolysis of municipal solid waste. Waste Manag. Res. 2015, 33, 267–274. [Google Scholar] [CrossRef]
  91. Li, Q.; Meng, A.; Jia, J.; Zhang, Y. Investigation of heavy metal partitioning influenced by flue gas moisture and chlorine content during waste incineration. J. Environ. Sci. 2010, 22, 760–768. [Google Scholar] [CrossRef]
  92. Furimsky, E. Characterization of trace element emissions from coal combustion by equilibrium calculations. Fuel Process. Technol. 2000, 63, 29–44. [Google Scholar] [CrossRef]
  93. Yu, J.; Sun, L.; Xiang, J.; Hu, S.; Su, S.; Qiu, J. Vaporization of heavy metals during thermal treatment of model solid waste in a fluidized bed incinerator. Chemosphere 2012, 86, 1122–1126. [Google Scholar] [CrossRef]
  94. Wang, R.; Zhao, Z.; Yin, Q.; Liu, J. Mineral transformation and emission behaviors of Cd, Cr, Ni, Pb and Zn during the co-combustion of dried waste activated sludge and lignite. Fuel 2017, 199, 578–586. [Google Scholar] [CrossRef]
  95. Zhang, J.; Han, C.-L.; Xu, Y.-Q. The release of the hazardous elements from coal in the initial stage of combustion process. Fuel Process. Technol. 2003, 84, 121–133. [Google Scholar] [CrossRef]
  96. Oboirien, B.O.; Thulari, V.; North, B. Enrichment of trace elements in bottom ash from coal oxy-combustion: Effect of coal types. Appl. Energy 2016, 177, 81–86. [Google Scholar] [CrossRef]
  97. Zheng, L.; Furimsky, E. Assessment of coal combustion in O2+CO2 by equilibrium calculations. Fuel Process. Technol. 2003, 81, 23–34. [Google Scholar] [CrossRef]
  98. Roy, B.; Bhattacharya, S. Oxy-fuel fluidized bed combustion using Victorian brown coal: An experimental investigation. Fuel Process. Technol. 2014, 117, 23–29. [Google Scholar] [CrossRef]
  99. Font, O.; Cordoba, P.; Leiva, C.; Romeo, L.; Bolea, I.; Guedea, I.; Moreno, N.; Querol, X.; Fernández, C.L.; Diez, L. Fate and abatement of mercury and other trace elements in a coal fluidised bed oxy combustion pilot plant. Fuel 2012, 95, 272–281. [Google Scholar] [CrossRef]
  100. Oboirien, B.O.; Thulari, V.; North, B. Major and trace elements in coal bottom ash at different oxy coal combustion conditions. Appl. Energy 2014, 129, 207–216. [Google Scholar] [CrossRef]
  101. Suriyawong, A.; Gamble, M.; Lee, M.-H.; Axelbaum, R.; Biswas, P. Submicrometer Particle Formation and Mercury Speciation Under O2−CO2Coal Combustion. Energy Fuels 2006, 20, 2357–2363. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Chen, Y.; Meng, A.; Li, Q.; Cheng, H. Experimental and thermodynamic investigation on transfer of cadmium influenced by sulfur and chlorine during municipal solid waste (MSW) incineration. J. Hazard. Mater. 2008, 153, 309–319. [Google Scholar] [CrossRef]
  103. Zhao, S.; Duan, Y.; Lu, J.; Liu, S.; Pudasainee, D.; Gupta, R.; Liu, M.; Lu, J. Enrichment characteristics, thermal stability and volatility of hazardous trace elements in fly ash from a coal-fired power plant. Fuel 2018, 225, 490–498. [Google Scholar] [CrossRef]
  104. Zhao, S.; Duan, Y.; Li, Y.; Liu, M.; Lu, J.; Ding, Y.; Gu, X.; Tao, J.; Du, M. Emission characteristic and transformation mechanism of hazardous trace elements in a coal-fired power plant. Fuel 2018, 214, 597–606. [Google Scholar] [CrossRef]
  105. Salgansky, E.A.; Podlesniy, D.N.; Tsvetkov, M.V.; Zaichenko, A.Y. Thermodynamic Estimating the Mass Transfer of Compounds of Rare Metals under Conditions of a Filtration Combustion Wave. Russ. J. Appl. Chem. 2020, 93, 1096–1101. [Google Scholar] [CrossRef]
  106. Liu, J.; Falcoz, Q.; Gauthier, D.; Flamant, G.; Zheng, C. Volatilization behavior of Cd and Zn based on continuous emission measurement of flue gas from laboratory-scale coal combustion. Chemosphere 2010, 80, 241–247. [Google Scholar] [CrossRef]
  107. Miller, B.B.; Kandiyoti, R.; Dugwell, D.R. Trace Element Emissions from Co-combustion of Secondary Fuels with Coal: A Comparison of Bench-Scale Experimental Data with Predictions of a Thermodynamic Equilibrium Model. Energy Fuels 2002, 16, 956–963. [Google Scholar] [CrossRef]
  108. Verhulst, A.D.; Buekens, A.; Spencer, P.J.; Eriksson, G. Thermodynamic Behavior of Metal Chlorides and Sulfates under the Conditions of Incineration Furnaces. Environ. Sci. Technol. 1996, 30, 50–56. [Google Scholar] [CrossRef]
Figure 1. Partitioning of Cd during coal combustion at fluidised-bed power station.
Figure 1. Partitioning of Cd during coal combustion at fluidised-bed power station.
Processes 08 01237 g001
Table 1. Cadmium levels in target materials.
Table 1. Cadmium levels in target materials.
MaterialCd Concentration (Range)References
Coal0.033–0.64 ppm[23]
Low rank coal (ash)0.24 (1.1) ppm[22]
High rank coal (ash)0.20 (1.2) ppm[22]
Municipal solid waste0–90 ppm[23]
Municipal solid waste incineration fly ash50–450 ppm[24]
Municipal wastewater sludge100 ppm[23]
Sewage sludge ash2.3–94 ppm[24]
Limestone0.01 ppm[25]
Urea0.03 ppm[25]
Table 2. Melting points (m.p.) and boiling points (b.p.) of Cd and its target compounds.
Table 2. Melting points (m.p.) and boiling points (b.p.) of Cd and its target compounds.
Compoundm.p./b.p.TemperatureReferences
MetalCd m.p.321 °C[38,39]
Cd b.p.767 °C[38,39]
OxideCdO m.p.1540 °C (sublimation)[39]
CdO b.p.Sublimation at 900 °C[38]
ChlorideCdCl2 m.p.564 °C[38,39,40]
CdCl2 b.p.960 °C[38,39,40,41]
SulfateCdSO4 m.p.1000 °C[38,40]
Table 3. Effect of adsorbents on Cd retention.
Table 3. Effect of adsorbents on Cd retention.
ExperimentEvaluatedAdsorbentResultsReferences
CdCl2, thermogravimetric reactor, 800 °CMetal adsorbedBauxite 74% *[45]
Alumina 55% *
Limestone 23% *
Emathlite12% *
Kaolinite11% *
Silica4% *
Synthetic solid waste (with Cd-nitrate), fluidized bed incinerator, 700 °C (900 °C)Cd adsorption efficiencyAl2O39% (4%) *[46]
Bauxite14% (9%) *
Kaolinite 5% (5%) *
Wood + Cd-acetate, fluidized-bed incinerator, 750 °CPercent captureBauxite68% *[47]
Zeolite 50% *
Lime40% *
Sorbent mixture75% *
Dried sewage sludge + 5% adsorbent, drop-tube furnace,
800 °C
Captured fractionKaolin44% *[48,49]
Zeolite 27% *
Limestone22% *
Scallop19% *
Mullite 15% *
Apatite14% *
Bauxite7% *
Silica4% *
Alumina2% *
Coal + kaolinite, electrically heated combustorRelative enrichment factor1100 °C0.4 *[50]
1200 °C2.1 *
1300 °C0.75 *
Coal + kaolinite, electrically heated combustor, 1100 °CConcentrationKaolinite0.38 ppm *[50]
Bauxite0.40 ppm *
CaO0.39 ppm *
Sewage sludge + additive (5:1), thermogravimetric analyzer, 1200 °CFixed ratio of CdNo additive25% *[51]
CaO20% *
Kaolin22% *
Coal + CaCO3 modified by 3 different additives, muffle furnace, 900 °CCd-capturing rateNa2CO322.83%[52]
K2CO357.37%
Al2(SO4)347.55%
Fluidized-bed incinerator, artificial solid waste (with Cd-nitrate), 900 °C% retention
(on silica sand)
no Na added5% *[53]
1.2% Na added46% *
* Rough estimation (data from diagram).
Table 4. Effect of Cl compounds on Cd volatility.
Table 4. Effect of Cl compounds on Cd volatility.
ExperimentParameter EvaluatedChlorine AddedResultsReferences
Bitum. coal + SRF *, ca. 1200 °CRelative enrichment factor in filter ash vs. cyclone ash (related to Al)-5–7[66]
1 and 2% NaCl8 and 10
2 and 4% PVC6 and 23
Bitum. coal + SRF *, ca. 1200 °CTEM-EDS analysis of filter ash aerosols from vaporization mode-<0.5%Cd/<0.5%Cl[66]
1% NaCl1.5%Cd/6%Cl
2% PVC1%Cd/1.5%Cl
Solid waste, 900 °C Volatilized fraction of Cd-47%[68]
1 and 3% Cl (NaCl)85 and 90%
1 and 3% Cl (PVC)90 and 95%
Wastewater sludge, 850 °CVolatilization rateNo NH4Cl (0.1% Cl)40%[69]
With NH4Cl (0.5% Cl)85%
Simulated MSW, 800 °CPercentage in fly ash and flue gas **No FeCl320% in fly ash ***,
80% in flue gas ***
[70]
With FeCl380% in fly ash ***,
20% in flue gas ***
MSW fly ash, 800 °CRemoval percentage of Cd from fly ash-50% ***[71]
100 g/kg Cl (NaCl)48% ***
100 g/kg Cl (MgCl2)90% ***
100 g/kg Cl (CaCl2)76% ***
* Solid recovered fuel; ** Percentage in bottom ash was within a few % (neglected); *** Rough estimation (from diagram).

Share and Cite

MDPI and ACS Style

Bartoňová, L.; Raclavská, H.; Čech, B.; Kucbel, M. Behavior of Cd during Coal Combustion: An Overview. Processes 2020, 8, 1237. https://doi.org/10.3390/pr8101237

AMA Style

Bartoňová L, Raclavská H, Čech B, Kucbel M. Behavior of Cd during Coal Combustion: An Overview. Processes. 2020; 8(10):1237. https://doi.org/10.3390/pr8101237

Chicago/Turabian Style

Bartoňová, Lucie, Helena Raclavská, Bohumír Čech, and Marek Kucbel. 2020. "Behavior of Cd during Coal Combustion: An Overview" Processes 8, no. 10: 1237. https://doi.org/10.3390/pr8101237

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop