Open access peer-reviewed chapter

Geopolymers and Alkali-Activated Materials for Wastewater Treatment Applications and Valorization of Industrial Side Streams

Written By

Tatiana Samarina, Esther Takaluoma and Outi Laatikainen

Submitted: 13 October 2020 Reviewed: 10 March 2021 Published: 28 March 2021

DOI: 10.5772/intechopen.97141

From the Edited Volume

Advances in Geopolymer-Zeolite Composites - Synthesis and Characterization

Edited by Petrică Vizureanu and Pavel Krivenko

Chapter metrics overview

535 Chapter Downloads

View Full Metrics

Abstract

The EU has the ambitious goal to transition from linear to circular economy. In circular economy, the old saying of “one’s waste is the other’s treasure” is being implemented. In this chapter, valorisation of industrial side streams, traditionally branded as waste, is discussed with respect to their applications as raw materials for new adsorptive products – geopolymers (GP) and alkali-activated materials (AAM) – as adsorbents in wastewater treatment. The chemical nature and structure of materials generally have great influence on GP/AAM adsorption capability. The approaches used for the raw materials preparation (chemical or physical) prior geopolymerization to increase the adsorption capacity of the final products will be discussed. Adsorption properties and performance of GPs/AAMs towards various contaminants are described, and the latest research on testing those materials as water remediation are reviewed. Special attention is paid to regeneration of exhausted materials and available resource recovery options that the regeneration approach opens. New forms of geopolymer adsorbent such as foams or core-shell structures are described and in the last part of the chapter, a short economic evaluation of resource recovery models is provided.

Keywords

  • adsorption
  • metal removal
  • nutrient recovery
  • geopolymer composite
  • wastewater treatment

1. Introduction

The United Nations have ratified 17 goals of sustainable development, of which responsible consumption and production is directly, while economy, innovative industry, infrastructure, and climate action are indirectly related to circular economy and the need of sustainable production [1]. Sustainability in the processing industries can be applied along the main value chain, e.g. from metal extraction to metal recycling, but can also be applied to the associated waste materials. Copper and iron mining alone are estimated to generate yearly about 5 bn tons of tailings [2], i.e. the fraction of the processed ore, after extraction of the valuable minerals. Finding a way to successfully reuse vast amount of this material and other waste sources is a great step towards circular economy.

While recycling an initial waste or side-stream, the material can be upcycled, meaning the newly derived product is of higher intrinsic value and properties or downcycled, where the new material has lower value. A prominent example for downcycling is the reuse of plastic bottles as fleece and carpet material. The material has less intrinsic value, because the carbon chains of the plastic polymer are shortened. Geopolymerization of industrial side streams is an upcycling process, as the geopolymer (GP), utilized as concrete and binder, has a higher value than the initial industrial by-products. To obtain upcycling, energy is put into the system, however, since tailings have usually been milled, the material has already undergone energy intensive steps and can therefore readily be used as starting blocks for geopolymerization. Utilizing tailings for upcycling into GPs, is therefore beneficially in terms of waste management, process energy, and emission of greenhouse gases, as the energy used in the beneficiation process is passed onwards into the geopolymerization process. Upcycling often requires further energy sources to achieve higher valuable material. In geopolymeration or alkaline activation this means the addition of chemicals, and in some occasions, such as analcime tailings, the addition of thermal energy [3].

In this chapter the valorization of high volume, inorganic side streams from mining, chemical industries, steel processing, and waste incineration into new adsorbents useable for water treatment is discussed. The purpose is to show how the material undergoes value change from side stream to potentially highly functional material.

As every tailing and every ash has a different chemical and mineral composition, tailoring of the properties of resulting adsorbents is possible by careful choosing of precursor materials. Aluminosilicates form the backbone of the geopolymer structure, but ion exchange, channel size, and physical properties are affected by the minerals used for geopolymerization [4, 5, 6]. Lastly, by controlling of the geopolymerization conditions, also the macroscopic structures can be developed by using various manufacturing methods from foaming to granulation.

The ultimate goal of using GPs/AAMs in water purification is to be able to recover valuable materials such as nutrients or battery chemical metals from contaminant-rich wastewater streams. In other words, the target is to use one industrial side stream to recovery of valuable material from another side stream or waste water in order to multiply circular economy potential.

Advertisement

2. Raw materials and preparation of GP/AAM materials for water treatment applications

This section summarizes different types of aluminosilicate precursors, occuring naturally or derived from industrial processes. Materials, which are currently abundant and/or urgent to dispose of, fall within the ambit of the section, but cover only water treatment applications not geopolymer production for construction industry, e.g. substitutes for Portland cement or as tailings’ covering.

2.1 Conventional and new sources of aluminosilicate precursors for GP/AAM preparation

Ashes.Fly ashes (FA) are prominent materials used as alkaline-activation binders. FAs are abundant yet complex materials, the composition of which is affected by co-incinerated material. Produced mostly by coal-fired electric and steam generating plants, coal FAs represent the greater part of generated FAs with the estimated flow of approximately 750 Mt. in 2015. The utilization rates of FAs differ greatly worldwide. Thus, rates in the developed countries equal 96%, 85%, 50% and 65% from produced streams for Japan, Germany, UK and USA, respectively [7]. For the developing countries with growing economics such India and China, the utilization rates are 38% and 45%, yet rate 66% for Asia in general are reported. Russian Federation and Africa implement FA as secondary resource less than 20% [7].

The composition of FA varies widely as it is derived initially from various primary sources: municipal waste/sludge co-incineration, different coal types, or subspecialized byproducts from industrial treatment plant (paper, forestry industry or agriculture). The combustion and cooling processes have profound impact on the characteristics of FA (particles size, shape, surface area, uniformity, etc.) as well as its composition and impurities’ inclusion.

Mainly, ASTM C 618 specification is applied to indicate the class of FA used for geopolymer preparation; however, a local/field or an unspecified labelling is also common. Coal FA (class F [8, 9] and C [10]) has been extensively considered as an aluminosilicate source for GP production, while the exploitation of biomass and co-incinerated FAs is less common [11, 12]. On the other hand, the utilization of these FAs particularly in the GP production for water treatment sector might be also beneficial. It would reduce the FA accumulation in landfills, and improve adsorbents’ LCA in comparison with metakaolin-based GPs.

Although FAs were studied as adsorptive materials previously [7, 13, 14], concerns on potential toxicity of impurities and convenience of use have encouraged to seek more suitable forms of FA-based materials for water treatment sector.

Municipal waste incineration bottom ash (IBA) has been traditionally considered as solid waste [15, 16]. IBA is mainly composed of Si, Al, Ca, and Na oxides, and could be classified as a hazardous or non-hazardous waste depending on the amount of toxic metal(oid)s. IBA’s main applications are in the engineering field as secondary materials in form of weathered bottom ash (after outdoor ageing for 2–3 months for pH stabilization). However, new applications of IBA have emerged in recent years as use as an aluminosilicate source for GPs/AAMs [17] including adsorbents [15, 18, 19, 20]. IBA by itself also was investigated as an adsorptive material for metal removal [21, 22]. In most of the studies, mixtures of IBA with various aluminosilicate precursors (BFS, FA, metakaolin) are used in order to obtain desired mechanical characteristics [23]. For the water treatment applications, IBA as raw material for GPs could have hidden benefits as the aluminum presented in it reacts with the alkaline activator and forms hydrogen gas, which leads to an increase in porosity [24]. Moreover, since a compressive strength of the resulting materials could be lower than for construction applications, the high porosity of materials and in situ stabilization of concomitant hazardous impurities via encapsulation could be attractive options [9, 19].

Pre-treatment of FAs and IBAs with various chemicals were suggested in order to reduce their toxicity and to meet the environment requirements of pristine materials or/and GPs/AAMs based on them [14, 25, 26, 27].

Steel industry waste.Blast furnace slag (BFS), dust, and sludge. BFS is another copious industrial nonmetallic by-product that is used widely as GP precursor in civil engineering [28]. Similar to ashes, steel industry wastes are mainly composed of Si and Al oxides, while Mg and Ca oxides could consist up to 35–60% of the material by weight [29]. Iron and sulfur are the major impurities in BFS, derived from the iron-smelting process. Ground granulated BFS has a high specific area due to the small particle size distribution, which makes it an excellent candidate as adsorptive material [13, 30]. However, in order to avoid the leaching of heavy metals from BFS during deployment of these materials in water remediation techniques, alkaline activation were suggested for entrapment/binding impurities within the GPs’ matrix [31, 32, 33, 34]. The presence of significant amounts of silica, aluminosilicates, and calcium-alumina-silicates in a pristine material makes the geopolymerization process rapid and effective resulting in rigid and enduring compositions. Thus, BFS was used to enhance the stability of metakaolin- and FA-based GPs [35].

Basic oxygen and electric arc furnace slag. These materials are sub-categorized, depending on the process of their formation. Both basic oxygen furnace slag (BOFS) and electric arc furnace slag (EAFS) are formed during the steelmaking process [36, 37]. These types of slags are similar in composition to BFS, except for their iron, manganese, chromium and sulfur contents, which are higher in BOFS and EAFS. EAFS was modified by alkali activation [38] and its adsorption properties towards copper were compared to raw EAFS. Significant amount of posnjakite were detected in the crystalline phase after adsorption of copper that could explain the drastically high removal efficiency of this AAM.

The accumulation of BOFS has become a significant issue due to its generation in large quantity, high disposal costs, and unsuitableness in cement industry due to high iron oxide content. Sarkar et al. adopted BOFS as a raw material for obtaining of GPs and investigated Ni2+ [39], Zn2+ [40], and F [41] removal. BOFS was used by Sithole et al. as a precursor for AAM preparation [42, 43]. In order to achieve highly porous structures for percolation column tests, a foaming agent (hydrogen peroxide) was added.

Red mud, silica fume and ore materials.Bauxite is a sedimentary rock, rich in aluminum oxide minerals and accompanied by kaolinite, quartz, and iron oxides. The amount of impurities varies depending on the place of origin. Bauxite itself has recently been tested in water treatment applications for purification of fluor- and arsenic-contaminated waters [44, 45]. However, even keener interest is observed in the valorization of bauxite residues, rich in iron and aluminum, as GPs/AAMs [46] and their water purification applications [47, 48].

Alumina manufactured through Bayer process from bauxite mainly goes to aluminum metal production. The rest (up to 10% of the whole production flow) is used as fillers in construction, in glass production, abrasive materials, and catalysts. Depending on the field of application, aluminum oxide could be called alumina, aloxide, aloxite, or alundum, and their respective waste materials would conform to these terms. Aloxite, for instance, is used as catalyst and/or catalyst support in organic chemistry due to its physicochemical stability and unique surface properties. As waste material, it could be accumulated in immense amounts from gas purification, decolorization, and catalytic processes as well as refining and desulfurization of petroleum oils and waxes. One of the ways of its valorization is in the design of eco-friendly GPs [49] and adsorbents for wastewater treatment [50, 51]. Thus, the addition of aloxite to analcime before geopolymerization showed an increase in the specific surface area and pore volume [50].

Silica fume is an amorphous form of silicon dioxide, collected as a by-product of the silicon and ferrosilicon alloy production. Historically, the main field of application is as pozzolanic material for high performance concrete of high strength and low porosity, though its applications for designing of foamed GPs and immobilization of cesium are reported [52, 53].

Mine tailings are mining and mineral wastes. The proper disposal of tailings has gotten under strong scrutiny for environmental preservation during the last decades. In many cases, the tailings are fine particles, containing silica together with iron oxides, alumina, and other minor minerals. This constitution makes tailings an excellent source of material for GPs. Iron ore tailings were mixed with FA to produce GP for copper removal [54], while gold mine tailings with Al2O3 additive were the source for GP production used for lead removal [51]. Prophyllite waste materials obtained from mine were converted to prophyllite GPs by Panda et al. [55], and tested as adsorbent for Co, Cd, Ni, and Pb removal from model solutions. Magnesite tailings from talcum mines show a reactivity dependent on the calcination temperature: light burnt (700 – 1000°C), hard burnt (1000 – 1500°C), and dead burnt (1500 – 2000°C). For use in water treatment and as a GP precursor, a light burnt magnesite in the form of periclase MgO is suitable.

Natural materials: zeolites, clays, sedimentary rocks. Kaolin is a rock rich in kaolinite, a clay mineral, and the source of production for the most widely used GP precursor called metakaolin. Metakaolin, a disordered, activated, and dehydroxylated form of kaolinite, is obtained through calcination of kaolinite over 600°C. The temperature of calcination has direct impact on reactivity of metakaolin, and as a consequence, on the crystallinity of the produced GPs [56]. However, raw kaolin/kaolinite were also used in GP production [57], yet for water treatment applications just recently [58]. Moreover, valorization of spent metakaolin could be beneficial and decline the cost of metakaolin-based GP [59]. Other clay materials have also been utilized recently for GP design [60]. Thus, bentonite clay calcined at 700–800°C were used by Maleki et al. [61] for obtaining a magnetic GP for heavy metal removal. Laterite clay-based GP were proposed by Ghani et al. [62] as a promising adsorptive material for Ni and Co removal. Laterite was activated at 900°C prior the geopolymerization. Volcanic tuff is another naturally available material with high porosity and with a high potential for ion-exchange. It was used in [63, 64] as an abundant yet low-cost material for GP production, and subsequent Zn removal from water.

Zeolites and zeolitic materials are well-known microporous materials. Found in nature or obtained through synthetic procedure, they are considered to be selective adsorbents [65, 66, 67], catalysts [68], carriers in biotreatment [69] due to their unique structure. Although, naturally occurred zeolites are readily available, they generally show lower surface area than synthetic ones.

Recently, much attention has also been paid on how zeolite could be synthesized from low-cost materials [70]. GP-zeolite composites and zeolite-like GPs are two different categories of adsorptive materials, which have recently attracted increased interest [71]. GP-zeolite composites are hybrid materials, unite the advantages of both constituents. The GP here serves as a durable support, while the zeolite provides a high surface area, porosity, and adsorption capacity. For instance, metakaolinite–zeolitic tuff GPs have been proposed in [72]. The report clearly showed the beneficial influence of the zeolitic tuff addition into a starting mixture on the microstructure and the adsorption potential of GPs. Andrejkovičová et al. [4] prepared metakaolin-based GPs blended with by 25, 50 and 75% of Nižný Hrabovec zeolite. It was shown that the zeolite particles are responsible for the higher amount of crystalline phases, producing a more compact and firm microstructure of blended GPs. The amount of blender has significant influence on the order of adsorbed metals and on the adsorption capacities of the formulations. Hayashi et al. [63] incorporated clinoptilolite into GPs though sol–gel protocol in order to further use of the resulting coatings for heavy metal ion adsorption.

It should be noted that zeolitic phase could be incorporated into GPs’ structures not only externally. Zeolite-like crystalline phases could be derived from synthesis routes through fusion method or even at moderate temperatures leading to zeolite-like GP structures. Javadian et al. [64] converted FA into a mesoporous aluminosilicate adsorbent through a fusion method at 600°C. Deng et al. showed that a hydrothermal synthesis of zeolite-like materials from IBA with higher crystallinity than through a fusion method is possible [73]. Similarly, Visa [74] converted FA into zeolite through a hydrothermal process. Rios et al. synthesized zeolite-like GPs from metakaoline at 100°C through the hydrothermal procedure [75]. Studies reported indicate that such materials have higher surface area and porosity than GPs/AAMs obtained through simple alkaline activation. Although the ultimate set of preferable conditions to form a GP instead of a zeolite are still under discussion, ratios Si:Al > 1.5 have been empirically established as providing more amorphous structures [60].

α-Analcime is a reject from spodumene refining at a Finnish lithium hydroxide plant, currently in piloting stage, and estimated to start the production in 2024, but is also found as a natural zeolite [3]. The small cavity size of analcime facilitates ion-exchange only for small mono- and divalent cations such as ammonium and Cu2+, and also K+, Ag+, Tl+, Rb+ (at elevated temperatures), and with low adsorption capacity. Raw analcime is inert to alkaline activation and analcime requires either chemical activation by 3–5 M H2SO4 or thermal activation above 700°C [3].

Not infrequently, industrial side streams cannot be used alone for geopolymerisation due to disharmonious Si/Al molar ratios. Therefore, by-products are commonly used as mixtures of aluminosilicate sources [76]. Table 1 summarizes the studies on different compositions of GPs/AAM that have been proposed for water and wastewater treatment applications. An afford was made to collect and match the precursors, synthetic protocol specificity, and distinctive characteristics resulting materials.

GP/AAMPrecursor/additivesPreparation method, prime oxide ratiosSurface Area/Pore Volume/Pore sizeType/form of GPRef.
MK-GP with TiO2MKHT27.21 m2/g 0.207cm3/g
2.19 nm
Bulk[77]
MK-GPMKSSM53.95 m2/g
0.061 mL/g
5.38 nm
Porous/Spheres, 2–4 mm[78]
MK-GPMKAA
Si/Al = 1.7
̶Bulk[79]
MK/Z-GPMK
Zeolitic tuff
AA
SiO2/Al2O3 = 1
̶Bulk/Discs[72]
MK-GPMKAA
Si/Al =
12.21 m2/g
0.037cm3/g
Bulk[80]
MK-GPMK
SDS 0.06 wt%
SSM
SiO2/Al2O3 = 1.6
53.95 m2/g
1.29 cm3/g
15 nm
Porous/spheres 2–4 mm[81]
MK/FA-GP2/3 MK
1/3 bioFA
(w/w)
AA
SiO2/Al2O3
H2O2
̶Porous/Monolith[11]
MK-GPMKAA foaming,
SiO2/Al2O3 = 5
̶Foam/Powder <100 μm[82]
MK-GP/alginate hybridMK
sodium alginate
AA + SSM
SiO2/Al2O3 = 1.6
16.2 m2/g
0.05 mL/g 11.5 nm
Bulk/Spheres 2–4 mm[83]
MK-GPMKAA
SiO2/Al2O3 = 3.2
39.24 m2/gBulk/Powder 150 μm[84]
MK-GP functionalized with CTABMK
silica fume
AA
CTAB
216 m2/g
0.22 cm3/g
Bulk/rubbles, 1.5 mm[85]
MK/FA-GPMK:FA
2:1 wt
AA7.9 m2/gPorous/discs[86]
MK-GPwaste MKAA
SiO2/Al2O3 = 1.5
̶Bulk/Powder, granules[59]
MK-GPMKAA
SiO2/Al2O3 = 3.2
39.24 m2/gBulk/Powder, 150 μm[87]
MK-GP activated with hull ashMK
BioFA
TiO2
SSM
SiO2/Al2O3 = 3.18*
̶Porous/Spheres 2–3 mm[88]
MK/FA-GPMK:FA 50:50 wt%SSM
SDS
̶Foam/
Spheres
[89]
MK-GP/alginate-chitosan hybridMK
alginate/chitosan
SSM
AA
0.5 wt% H2O2,
1.5 wt% SDS
230 m2/g
0.99 mL/g
35 μm
Porous/Spheres[90]
MK/FA-GPMK
FA class C
60:40 wt%
AA
SiO2/Al2O3 = 2.7
̶Bulk/Powder
63–125 μm
[91]
MK/BFS-GPMK
BFS
60:40 wt%
AA
SiO2/Al2O3 = 3.1
̶Bulk/Powder
63–125 μm
[91]
MK-GPMKAA
SiO2/Al2O3 = 2.31**
̶Bulk/Powder, 200 μm[92]
MK-GP/
coal gangue
hybrid
MK
gangue
50/50 wt%
AA
SiO2/Al2O3 = 4.0
26.41 m2/g
0.330 cm3/g
Bulk[93]
MK-GPMKAA
SiO2/Al2O3 = 2–8
̶Bulk/Pervious[94]
MK-GPWaste MK Aluminum scrap recycling waste
1:1 (w/w)
AA
SiO2/Al2O3 = 1.25**
15.95 m2/gBulk/Granules 4–11.2 mm[95]
MK-GPMKAA8.16 m2/g
0.021 cm3/g
10.5 nm
Bulk/
Granules 0.5 mm
[96]
MK-GP
functionalized with HDTMABr
MKAA
HDTMABr
̶Bulk/Powder, 53 μm[97]
MK-GP
MK/aloxid
ANA-GP
ANA/aloxid-GP
MK analcime
aluminum oxide
AA**
MK-GP
SiO2/Al2O3 = 3.96
MK/aloxid GP
SiO2/Al2O3 = 2.13
ANA-GP
SiO2/Al2O3 = 7.01
ANA/aloxid-GP
SiO2/Al2O3 = 3.60
MK-GP
19.97 m2/g
0.131 cm3/g
26.24 nm MK/aloxid GP
6.36 m2/g
0.036 cm3/g
23.18 nm ANA-GP
0.69 m2/g
0.003 cm3/g
21.69 nm ANA/aloxid-GP
38.29 m2/g
0.125 cm3/g
13.07 nm
Bulk/Powder, 63–125 μm[50]
MK/Z-GP25% MK
75% zeolite
AA
SiO2/Al2O3 = 1
57.5 m2/gBulk[4]
MK-GP functionalized CTABCalcinated halloysite clayPrecipitation
SiO2/Al2O3 = 2.91
CTAB/Cu2O/TiO2
34.8 m2/g
29.7 nm
Bulk[98]
MK-GP
functionalized with CTAB
MKAA
CTAB
26.45 m2/g
0.121 cm3/g
9.12 nm
Bulk/Powder, 125 μm[99]
MK-GP,
magnetic
hybride
MK
Magnetite 5 wt%
AA
SiO2/Al2O3 = 4.55* H2O2
19.5 m2/g
0.045cm3/g
10.4 nm
Porous[100]
MK/Silica-GP
functionalized with Cr
MK
silica fume 9:1 (w/w)
AA
SiO2/Al2O3 = 1.90**
30 nmBulk/Membrane[101]
MK-GPMKAA
SiO2/Al2O3 = 2.14**
̶Bulk/Powder, 355 μm[102]
MK/FA-GPMK
FA
AA
SiO2/Al2O3 = 2.45**
27 m2/gBulk/rubbl 1.0–0.3 mm[103]
MK-GPMKAA
SiO2/Al2O3 = 4.0
21 m2/g
1252 mm3/g
0.32 μm
Bulk/Granules,
3 mm
[104]
MK-GP
alginate hybrid
MK
sodium
alginate
TiO2
SSM
SiO2:Al2O3 = 4
20 m2/g
714 mm3/g
0.11 μm
Bulk/Spheres,
2 mm
[105]
MK-GP/
magnetic hybrid
MKAA
SiO2/Al2O3 = 4.55
H2O2
42.92 m2/g
0.052cm3/g
4.88 nm
Porous[106]
MK-GP/LECAMK
LECA support
AA
SiO2/Al2O3 = 1.5
̶Bulk/Granules, 4–8 mm[107]
MK/Biochar-GPMK
Biochar
AA
H2O2
37.46 m2/gFoam/membrane[108]
MK-GP
functionalized with K4Fe(CN)6
MKAA
SiO2/Al2O3 = 3.60
H2O2
35 m2/g
55 cm3/g
Foam[109]
MK-GP/graphene oxide hybridMK graphene oxide 10 wt%AA
SiO2/Al2O3 = 0.45
̶Bulk/
Particles,
< 0.5 mm
[110]
FA-GPFly ashAA
SiO2/Al2O3 = 2.03
̶Bulk[111]
FA-GPFly ash, 75 μmFM
SiO2/Al2O3 = 1.98*
8.22 m2/g
2.9 nm
Bulk/Powder[64]
FA-GP Iron-enrichedCalcinated FA,
< 70 μm
FM
SiO2/Al2O3 = 1.00
Fe2O3/Al2O3 = 0.151
̶Bulk[112]
FA-GP modified with ironCoal fly ashAA
SiO2/Al2O3 = 1.43*
162.38 m2/g
0.126 cm3/g
3.90 nm
Bulk/Powder[113]
FA-GPFly ashAA
SiO2/Al2O3 = 4.61
̶Bulk/Powder, 71–90 μm[9]
FA/IOT -GPFly ash
IOT
70:30 (w/w)
AA
H2O2
6 nm - 360 μmPorous/Cubes[54]
FA-GPFly ashAA
SiO2/Al2O3 = 1.12**
20.48 m2/g
19.62 nm
0.070 cm3/g
Bulk/Powder, 74 μm[114]
FA/Z-GPFly ash
Fajustite
HT
SiO2/Al2O3 = 0.69**
174.35 m2/g
0.14 cm3/g
9.69 nm
Bulk/Powder, 74 μm[114]
FA-GPFly ash CAA
SiO2/Al2O3 = 3
̶Bulk/Powder[115]
FA-GP/LECAFly ash C LECA supportAA
SiO2/Al2O3 = 1.5
̶Bulk/Granules, 4–8 mm[107]
FA-GPFly ashAA
SiO2/Al2O3 = 5.36
̶Bulk/Powder, 71–90 μm[116]
FA-GPFly ashHT̶Bulk[117]
FA/Z-GPFly ash
BFS
4:1 (w/w)
HT
SiO2/Al2O3 = 3.49*
̶Bulk/Powder[118]
Fly ash/
iron oxide hybrid
Fly ash
Fe2O3 5 wt%
AA
SiO2/Al2O3 = 3.30**
60.75 m2/gBulk/Powder, 50 μm[119]
FA-GP/Graphene hybridFly ash graphene (1 wt%)AA
SiO2/Al2O3 = 3.41**
20.41 m2/g
0.047 mL/g
9.73 nm
Bulk[120]
FA/BFS-GPFly ash BFSHT
SiO2/Al2O3 = 3.23*
76.6 m2/g
0.24 cm3/g
12.5 nm
Bulk[35]
FA-GPBoiler fly ash
< 80 mesh
HT
SiO2/Al2O3 = 2.75**
27.51 m2/g
0.032 mL/g
Bulk[121]
FA-GPFly ashAA
SiO2/Al2O3 = 3
29 m2/g
0.134 cm3/g
Bulk/Powder, < 74 μm[122]
FA-GP/Polyethersulfone hybridFly ashAA
SiO2/Al2O3 = 3.05
168.3 m2/gBulk/Powder, 150 μm[123]
FA/Z-GPCalcinated fly ashSiO2/Al2O3 = 1.61**̶Bulk/Powder[124]
FA-GPFly ashAA131.4 m2/gBulk/Powder, <105 μm[125]
FA-GPCoal Fly ashFM
SiO2/Al2O3 = 1.25*
93.8 m2/g
0.62 cm3/g
Bulk[8]
FA-GPFly ash
silica
AA
Si/Al = 2.2
31.87 m2/g
0.12 cm3/g
15.45 nm
Bulk/Powder, 125–212 μm[126]
FA-GPFly ash F,
≤ 177 μm
AA
SiO2/Al2O3 = 2.97**
30 m2/g
0.076 cm3g
Bulk[127]
FA/analcime-GPFly ash
analcime
AA
SiO2/Al2O3 = 2.10**
̶Bulk/Membrane[128]
FA-GPFly ashAA
SiO2/Al2O3 = 5.42
35.97 m2/g
124 cm3/kg
9 nm
Bulk/Powder, 150 μm[31]
FA-GPRice husk ash, waste alum cansHT
SiO2/Al2O3 = 1.82*
36.15 m2/g
0.097 mL/g
5.4 nm
Bulk[129]
FA-GPFly ash class C and FAA
SiO2/Al2O3 = 6.6**
(class C)
SiO2/Al2O3 = 10.9** (class F)
2463.64 mm2/g̶[130]
FA/MK-GPbioFA
MK
70:30 (w/w)
AA
aluminum powder, anionic surfactant
46.3 m2/gFoam/Membrane[131]
IBA-GP/Graphene hybridBottom ash grapheneAA
0.15 wt% Mn2+
19.5 wt% CuO
29.28 m2/g
0.1078 mL/g
14.77 nm
Bulk/
Particles, 0.180–0.315 mm
[132]
BFS –GPBFSAA
SiO2/Al2O3 = 4.40*
64.5 m2/g
0.095 cm3/g
5.93 nm
Bulk/Powder, 63–125 μm[32]
BFS –GPBFSAA
SiO2/Al2O3 = 3.2
̶Bulk/Powder, 63–125 μm[115]
BFS –GP/graphene hybridBFS graphene 0.01 wt%AA SiO2/Al2O3 = 2.61*146.17 m2/g
0.161 mL/g
4.40 nm
Bulk/Powder, 250-315 μm[133]
BFS –GP/barium modifiedBFSAA
SiO2/Al2O3 = 4.00**
63.1 m2/g
0.070 cm3/g
Bulk/Powder, 63–125 μm[134]
BOFS-GPBOFSAA
SiO2/Al2O3 = 11.5**
30.84 m2/g
0.091 cm3/g
11.8 nm
Bulk/Particles, ∼0.1 mm[39, 40]
Slag-based GPSlagSSM
SiO2/Al2O3 = 4.02*
0.3 wt% SDS
100.9 m2/g
7 nm
Porous/Spheres, d ≈ 100 μm[33]
Silicomanganese slag-GPSilicomanganese slag
(NH4)6Mo7O24·4H2O
AA
SiO2/Al2O3 = 1.44**
51.79 m2/g
0.192 mL/g
10.30 nm
Bulk/Particles, 0.16–0.315 mm[135]
BOFS-GP modified with Ni(II) or Zn(II)BOFS-GPAAZn/LDS-GP
58.14 m2/g
Ni/LDS-GP
53.42 m2/g
LDS-GP
30.84 m2/g
Porous/Powder, ∼0.1 mm[41]
Slag-based GPSlagSSM
SiO2/Al2O3 = 3.08*
87.74 m2/gBulk/Spheres[34]
Slag-based GP/
Fe2O3-hybride
SlagSSM
SiO2/Al2O3 = 70.65**
Fe2O3/Al2O3 = 188
Fe2O3/SiO2 = 2.66
233.8 m2/gBulk/Microspheres, 75–300 μm[136]
Steel slag/fly ash/analcime-GPSteel slag fly ashHT
SiO2/Al2O3 = 2.01**
27.25 m2/g
0.050 cm3/g
8.12 nm
Bulk[137]
BFS –GPBFSAA
SiO2/Al2O3 = 5.26
23.56 m2/g
73 cm3/kg
7.8 nm
Bulk/Powder, 150 μm[31]
EAFS-GPelectric arc furnace slagAA
SiO2/Al2O3 = 2.02*
6.5 m2/g
0.014 cm3/g
8.7 nm
Bulk/Powder[38]
BOFS-GPBasic Oxygen furnace slagAA
H2O2
̶Porous[43]
Slag-GP/CeO loadedSlagSSM
SiO2/Al2O3 = 3.31**
186.40 m2/g
0.352 cm3/g
7.56 nm
Bulk/Sphere, 75–300 μm[138]
Clay-based GPKaolinFM
SiO2/Al2O3 = 1.88**
51.3 m2/g
0.324 cm3/g
25.25 nm
Bulk/Powder[58]
clay/gangue microsphere -GPKaolin
coal gangue 50/50 wt%
AA
SiO2/Al2O3 = 4.0
39.74 m2/g
52.00 nm
Bulk/rubbles, 0.45–0.15 mm[139]
Clay-GP/Fe3O4 hybrideCalcined bentonite clayAA2.32 m2/g
0.008 cm3/g
13.76 nm
Bulk/Powder[61]
Clay-GPLateritic clay, 58 μmAA17.441 m2/g
0.005 cm3/g
1,4 nm
Bulk/Powder, 58 μm[62]
Natural tuff-GPVolcanic tuffAA
SiO2/Al2O3 = 3.74**
̶Bulk/Powder, < 200 μm[140]
Alumino silicate-GPAlumino silicate powderAA
SiO2/Al2O3 = 4
50.1 m2/g
0.36 cm3/g
0.04 μm
Bulk/Monoliths or granules[141]
Synthetic GPChemosynthetic Al2O3-SiO2 powderSSM
SiO2/Al2O3 = 2
̶Bulk/Spheres[142]
Chitosan modified geopolymerAluminum salt and silica solution,
chitosan
Precipitation
SiO2/Al2O3 = 3.06**
̶Bulk/Powder[143]
OTB-GPPyrophyllite mine waste samplesAA
SiO2/Al2O3 = 2.39**
̶Bulk/Powder, <45 μm[55]
OTB-GPGold mine wasteFM74.92 m2/gBulk[144]
OTB-GPGold mine tailings
Al2O3
FM74.916 m2/gBulk[51]
Municipal solid waste-GPSludgesFM
SiO2/Al2O3 = 3.12**
0.496 m2/g
9.98 nm
Bulk[145]
Municipal solid waste-GPMunicipal solid waste biocharAA6.5 m2/g
-
-
Bulk[146]
Dolochar ash based geopolymerDolochar
< 100 mesh
AA
SiO2/Al2O3 = 4.97**
49.91 m2/g
0.087 cm3/g
8,9 nm
Bulk/Particles, ≈ 0.1 mm[147]

Table 1.

GP/AAMs compositions for water and wastewater treatment reported in literature.

Calculated using the amounts of raw materials in the slurry.


Calculated using the XRF of product.


Materials: GP – geopolymer; MK – Metakaolin; FA – fly ash; IBA – incineration bottom ash; BFS – blast furnace slag; BOFS – basic oxygen furnace slag; EAFS – electric arc furnace slag; Z – zeolite; IOT – iron ore tailing; OTB-ore/tailings based; Procedure: AA – alkaline activation at moderate temperature; SSM – suspension-solidification method; FM-fusion method; HT- Hydrothermal method; Additives: SDS – sodium dodecyl sulfate; CTAB – cetyltrimethylammonium bromide; HDTMABr -hexadecyltrimethylammonium bromide.

2.2 Forms and manufacturing techniques of GPs/AAMs for water treatment

Originally, a basic composition applied for manufacturing GP/AAM adsorbents consisted of an alumosilicate precursor, an alkali, and an additional source of silicate in a form of water glass. Initially, both sodium and potassium forms of alkaline activators were used to induce geopolymerization. In the vast majority of the research reviewed, sodium alkaline and water glass are used in the activation process. It was shown by Bakharev that dissolution rates of the minerals was higher when a sodium form is used [148]. Luukkonen et al. [149] found that adsorption characteristics of metakaolin-based GP prepared with NaOH is better than with KOH in case of ammonium removal. An in-depth discussion of G chemistry and vivid explanations could be found in the latest reviews [57, 150, 151].

Forms and manufacturing techniques of GPs/AAMs for water treatment application are emerging and evolving constantly. In the first instance, powdered forms were mostly used for gaining of adsorption characteristics of materials. At first, GPs/AAMs were manufactured in bulk forms to be crashed to powder or rubbles after curing procedures. However, these materials have relatively low porosity, and addition of foaming agents were appealing for increasing surface area, pore volume, and porosity. In Table 1 forms of GPs/AAMs reported and specific additives listed along aside with their synthetic procedures and properties gained. For bulk samples, species are usually sealed with plastic films to prevent moisture evaporation since the presence of water increases porosity. Curing and aging are usually carried out at temperatures ranging from 20 to 105°C. Commonly, an industrial application of powdered forms requires instance pressure filtration and an additional separation step after adsorbent exhaustion. Both these processes increase a cost of treatment, its complexity, limiting a regeneration ability causing sludge accumulation.

Granular forms of GPs/AAMs are more preferable for large-scale applications. FA-based GPs supported on inert substrate were proposed with the aim of overcome these limitations [107]. A simple technique similar to the conventional GP preparation was applied to design floatable light granules for phosphorous removal. Moreover, spherical granules could be produced by in situ geopolymerization during granulation by a high-sheer granulator, where a liquid alkaline activator acts as a binding liquid [59, 152]. A granule size distribution is a function of a liquid to solid ratio, granulation time, and a rotation speed. While the amount of liquid required depends on the wetting behavior and particle size of the precursors, a good starting point is L/S of 1/3. The production method is easy to upscale that makes feasible large-scale water treatment applications with granular GPs. A more complicated procedure enables the development of granular adsorbents through a suspension and solidification method, resulting in microspheres (<100 μm) [33, 34] or highly porous GPs [93]. Composites of GPs with biopolymers, for instance alginate that possesses the ability to solidification in presence of calcium ions, were also obtained in granulated form [83, 93].

Porous GP/AAM adsorbents contain relatively high volume of voids or pores. The pore sizes usually range from nanometers up to millimeters with the total pore volume ranging from 30 to 90% [153]. Direct foaming, either chemically or mechanically, is seen as the most widely applied foaming approach. The common additives that have been used in the direct foaming methods are hydrogen peroxide [11, 33, 43, 54, 78, 93], Al [131, 154] or Si [155] powders along aside with stabilization agents such surfactants or oils.

Pervious GP/AAM is another promising form of an adsorptive material for water purification. Thus, AAM-based membranes with potential to remove alkaline earth metals [156], and nickel [157] have been reported. Development of porous/pervious GPs/AAMs led to variety of fabricated adsorptive forms such as monoliths, membranes, granules, and self-supported filters. That is opening the versatility of approaches, conventional in water and wastewater treatment practice, yet challenging if the powdered forms used. Separation, regeneration, and surface modification are no longer restricted by the form of production. Porous/pervious materials can be used directly in packed bed adsorbers, and be easily regenerated or retrieved after adsorbent saturation with target substances or contaminants.

Advertisement

3. Properties and performance of GPs/AAMs towards various contaminants

Despite the fact that the first identification of GPs as unconventional construction materials was in 1979 [158], broader applications of GPs/AAMs started in late 90s. Although GPs/AAMs are to be considered by some authors as an economic alternative to zeolites or activated carbons for water purification, the lack of real cases reported is obvious. To urge commercial importance, GP/AAM adsorbents should be readily available, economically feasible, steady in characteristics, and easily regenerated. Several comprehensive reviews on the GP/AAM materials for the water treatment sector have been published just recently [57, 150, 151, 153]. Therefore, in this section the bright and promising works will be highlighted as well as challenges and trends for future studies revealed.

GPs/AAMs for metal(oid)s removal. The adsorption characteristics of individual species and particular conditions of adsorption could be found elsewhere [57, 150, 151, 153]. Here, we would like to emphasize some challenges and gaps, which might be addressed in future studies. There are only several articles discussing selectivity of adsorption on GPs/AAMs alongside the matrix effects. In most of the studies pure mono-element aqueous solutions were implied, and the adsorption characteristics for individual substances- without possible influence of matrix macro-elements have been established. However, GPs/AAMs that are considered as a replacement of zeolites have to demonstrate selectivity under complex matrices in order to have opportunities to promote the implementations in various industrial applications.

In order to obtain adequate adsorption parameters, an excessive alkaline residue in GP/AAM should be washed out properly (pH 7 ± 0.5 within 24 h required) [159]. Otherwise, the increment of pH of aqueous solutions containing heavy metals will favor the hydroxide precipitation process, leading to wrong result interpretation. For porous GPs, washing away the excessive alkalis resulted in the increment of total porosity [11], which led to better performance. Moreover, excessive alkalis were used intentionally to neutralize AMD [42] and remove metal ions. However, a strict protocol must be followed to characterize newly designed materials.

Selective adsorption is relies on several factors such as a metal ion activity, hydration radius and free energy of hydration, and a pore size distribution of GP.

Geopolymerisation by itself could lead to the formation of new ion-exchange sites at the GP surface, but additives in composite formulations could have even higher influence the adsorption characteristics.

An ionic exchange reaction between the heavy metal ions and sodium ions has resulted in heavy metal removal by the metakaolin GP [159]. The adsorption selectivity of heavy metal ions by the GPs at pH 4 in multi-component solution was in the following order: Pb2+ > Cd2+ > Cu2+ > Cr3+, while qe [mg/g]: 100 > 76 > 55 > 10. The order of adsorption was in accordance with the hydrated radius and free energy of hydration for selected ions. However, the free energy of hydration and the activity for Cr3+ are all higher compared to those of other metals, though its adsorption rate does not correspond to the assumed order. The selectivity towards Cr3+ was be explained through its ionic status. When the pH exceeded 4, Cr3+ transforms to Cr(OH)2+, which might lead to its lower adsorption ability. It is also noted that at lower pH, the balancing ions present on the GP surface tend to be replaced by the hydrogen ions instead of the metal ions that lead to lower capacity at acidic pH.

Lopez et al. [5] investigated the selectivity of metakaolin-based GPs in multicomponent solutions (Pb2+, Cu2+, Cd2+, Ni2+, Zn2+ and Cs+). For a composition with Si/Al ratio 2, the best capacities and selectivity towards Pb2+ and Cs+ were observed. The adsorption selectivity for the mixture of metal ions was in the following order Cs+ > Pb2+ > Cu2+ > Zn2+ > Ni2+ > Cd2+, while qm [mg/g]: 43 > 35 > 15 > 3 > 1 > 2. The adsorption capacity for individual elements were higher: 57 mg Pb2+/g > 52 mg Cs+/g > 46 mg Cu2+/g > 14 mg Cd2+/g > 9 mg Zn2+/g > 4 mg Ni2+/g. Moreover, the effect of solution salinity (NaCl, 5% and 10%, wt) was studied, and no considerable effect on the adsorption order of metal ions or GP capacity in multi-composition solution was found. The authors presumed the existence of at least two types of binding sites with different affinities toward the metal ions to explain such a tolerance.

Selectivity of GP composites with zeolite filler was studied by Andrejkovičová et al. [4]. The highest adsorption was observed for Pb2+ for all the GPs obtained, while an adsorption order was as follows: Pb2+ > Cd2+ > Zn2+ > Cu2+ > Cr3+. The adsorption of Cu2+ and Cr3+ increased as the amount of metakaolin in the GP increased, whereas the composite with 25% zeolite doping had higher adsorption characteristics towards Pb2+, Cd2+ and Zn2+. GPs prepared from zeolitic tuff and kaolinitic soil by El-Eswed et al. [160] showed totally different order of adsorption: Cu2+ > Pb2+ > Ni2+ > Cd2+ > Zn2+. Moreover, the adsorption order strongly depended on the GP composition, although Cu2+ and Pb2+ adsorption has always prevailed.

The ability of BFS- and metakaolin-based GPs to remove Ni2+ and metalloids (As and Sb) in form of oxyanions was shown in [32]. Both adsorbents completely removed Ni2+ that most likely was associated with precipitation of its hydroxides on the GPs, while both metalloid oxyanions were adsorbed by BFS-GP equally. Another remarkable merit is that the adsorption capacities were obtained with real matrixes (spiked mine effluents), and were 4.42 mg/g, 0.52 mg/g, and 0.34 mg/g for Ni2+, As3+, and Sb3+, respectively. It is specified by the authors that the low capacities could be a result of competition of some matrix ions (Sr, Ca, Mg, Mn) with the target ions for binding sites.

Researches with increasing frequency pay attention to this problem and try to demonstrate the removal efficiencies with real samples. Removal of Ca2+ and Mg2+ from intact groundwater was examined in [58] on kaolin-based GP. With adsorbent dose of 1 g/L, the removal rate were 37.5% and 16.2% for Ca2+ and Mg2+, respectively. Metakaolin-based GP was tested by Kara et al. [87] for Mn2+ and Co2+ removal from real wastewater. The removal rates in real wastewater decreased from 97.5% to 53.01% and 94.6% to 39.12% for Co2+ and Mn2+, respectively. The results demonstrated that the adsorption performance affected negatively by the coexistence of some other cations and/or anions in the adsorption medium. Bentonite-based GPs were used for heavy metals removal from synthetic wastewater [61]. Porous biomass FA-based GPs were used in [129] for simultaneous removal of heavy metals from wastewater samples. Mixed FA/metakaoline-based GPs were used in [103] for Cu2+ removal from real wastewater. In the showcase, the adsorption capacity of GPs towards Cu2+ decreased by 27% as compared to synthetic samples. Sithole et al. treated acidic industrial effluents by FA/BOFS-based GPs [42, 43]. New GPs containing hollow gangue microsphere were applied for Zn2+ removal from smelting plant wastewater in [93]. At an adsorbent dose of 30 g/L, a complete Zn removal was observed. The distinctive aspect of the reported cases was that a complex composition of treated solutions is likely to decrease substantially capacity of the GP. Thus, the adsorption capacities obtained for the ideal laboratory conditions should be primary used as the guiding not decision-making parameters.

GPs/AAMs for removal of other inorganic ions. Besides metal(oid)s, GPs/AAMs were examined for removal of ammonium and various anions. Luukkonen et al. [149, 152, 161] showed potential of metakaolin-based GPs to remove ammonium. The optimized GP composition was proposed and manufactured in both powder and granular forms. The efficiency of removal was demonstrated in municipal wastewaters (primary and secondary effluents) as well as landfill leachates. Metakaolin-based GPs prepared from commercial and waste metakaolin were able to effectively remove ammonium from synthetic and wastewater samples [59]. In fact, GPs prepared from paper mill fiber sludge showed better selectivity in the presence of competing ions under real matrix conditions. Bai and Colombo prepared metakaolin-based GP foams in the form of monolithic porous filters [162, 163]. The filter was able to remove up to 95.3% of ammonium from runoff waters at the initial concentration of 3 mg/L.

The removal of phosphorus was attempted in [10] with a pervious FA-based GP. The removal rate increased with the increase of pH. Up to 85% of phosphorus were removed from a treated wastewater. Simultaneous removal of ammonium and phosphate by composite metakaolin/BFS-based GPs was demonstrated in [91]. Phosphate removal was enhanced in presence of ammonium. At slightly alkaline conditions (pH 7–8), the removal rate towards phosphate ions was relatively high (>86%), whereas the ammonium removal up to 35% was also achieved. FA-, BFS- and fiber sludge GPs were investigated as promising adsorbents for phosphorous removal from diluted solutions. The capacities at initial phosphate concentration of 100 mg/L are 26 mg PO4/g for BFS-GP, 36 mg PO4/g for FAF-GP, and 43 mg PO4/g for FSHCa-GP [115].

Sulfate ions were removed by barium-modified BFS-based GPs [134]. Adsorption capacities were 91.1 and 119.0 mg SO4/g for model solution and mine effluent, respectively. The surface complexation or precipitation of barium sulfate were suggested as probable removal mechanisms.

Removal of halides by GPs/AAMs is an emerging topic. For this end, composite or functionalized materials are designed. Removal of F ̶ was demonstrated by slag-based GP microspheres modified with CeO [138], Fe2O3 [136], and bivalent metallic species [41] with capacities towards the contaminant 127.7 mg/g, 59.8 mg/g, and 60 mg/g (zinc impregnated BOFS-GP), respectively. A metakaolin-based GP functionalized by surfactant was developed for efficient removal of radioactive iodide [97]. High concentrations of competitive anions had limited influence on the adsorption process.

GPs/AAMs for removal of organic substances. In fact, GPs contained residual metal oxides could have potential catalytic performance. Thus, GPs based on industrial by-products such as FA, BFS, or their mixtures with silica fume and aloxite demonstrated catalytic activity under visible light irradiation. The descriptive list of organic substances removed by GPs could be found in reviews [57, 150, 151, 153]. Manly, cationic and neutral dyes were investigated as targets in recent studies, although removal of fecal coliforms [10], volatile organic compound [77], and tetracycline [164] was reported.

Oxidative degradation or photodegradation after adsorption have been specified by authors as primary mechanisms of organic pollutants’ removal. Although conventional GPs have been reported for these purposes [86, 89, 104, 126, 139], they would rather have had low adsorption/degradation characteristics. Hybrid or composite materials were proposed to improve the removal efficiency of organic pollutants. Thus, graphene [120, 132, 133, 165], TiO2 [88, 98, 105], CdS [142], various metal oxides [101, 106, 135] were introduced in GP matrix in order to enhance degradation abilities of resulting materials.

Advertisement

4. Regeneration of GPs/AAMs and further resource recovery options

In last a few decades, significant improvements were made in both efficiency and economy in removal of metal(oid)s and other substances by adsorbents. Nevertheless, regeneration and recycling of used adsorbents, or recovery of the removed species from the desorbing agents are still rarely reported. For regeneration and reuse of GPs/AAMs, various possible regenerating agents such as acids, alkalis and chelating agents could be used. Only a few of the reported studies were focused on recovery of adsorbed (from saturated adsorbents) and desorbed (from regenerating agents) metals [11, 87, 96, 131]. However, for industrial application and success completion of new GP/AAM adsorbents on the market, research studies on number of adsorption–desorption cycles are in high demand. Moreover, revenues gathered from resource recovery options will have a decisive role in further technology implementation.

The regeneration of metakaolin-based GP by sodium chloride under alkaline conditions after ammonium adsorption for the first time were demonstrated in [152]. Three adsorption–desorption cycles were carried out with a steady removal efficiency. Sodium chloride and sulfate, potassium sulfate and phosphate were studied in [59] as regenerating agents for saturated metakaolin-based GPs. Sodium sulfate showed better results during five cycles under continuous sorption–desorption experiment, only 34% of an initial overall capacity of the GP were lost. Sodium chloride regenerant was also efficient, but only 55% of ammonium could be removed after 5th desorption cycle. The same adsorbents were used to test a nitrogen recovery option in a laboratory-scale demonstration setup [166]. The layout consisted of an adsorption/desorption unit and Liqui-Cel® membrane. A liquid phase obtained during adsorbent regeneration was purified in the membrane contactor in order to recover ammonium nitrogen as ammonium sulfate or phosphate. The purified regeneration solution was used repeatedly for further adsorbent regeneration. Several regeneration-purification cycles were conducted to estimate system sustainability and chemical consumption demand. Operational conditions of a membrane process such as shellside and lumenside feed flows, temperature, and pH were adjusted to gain maximal capacity of the setup. One membrane contactor (2.5 × 8-inch Liqui-Cel) was used under following operational conditions: 100 L/h shellside and 60 L/h lumenside feed flows, 40°C working temperature, pH ≥ 10. Technical sulfuric or phosphoric acids, up to 5%, were used as lumenside phases. The concentration of ammonium-content salt in a resulting received phase were 17% and 22% for phosphate and sulfate salt, respectively.

Metal recovery from GPs/AMMs via ion-exchange mechanism can only take place if physical adsorption occurred and the pH was low enough to prevent precipitation of metal hydroxide during adsorption process. Acids of over 0.1 M strength affect the structure of the GPs, and while metals are regenerated by acid washing, the reuse of adsorbents are diminished both in batch [11] as also in continuous mode [87, 167] experiments. Mild acid washing with 0.01 M H2SO4 or HNO3 removed metals from GPs efficiently in short time (1–2 h). It has also been shown that the adsorption capacity after mild acid washing could increase [131], which could be explained by exchange of Na+ with easier replaceable H+ cations. Selective desorption of copper has been observed by ammonia. A linear desorption ability with respect to ammonia concentration was observed, and complete desorption being possible by 10% ammonia solution [50, 61].

Sequential desorption tests of Cd2+ have been conducted on a loaded metakaolin GP, establishing the percentages of physically adsorbed, ion-exchangeable, EDTA extractable, and residual forms of metal [96]. The authors showed that physical adsorption is negligible, and ion-exchange with MgCl2 constituted to only 2–8% of adsorbed Cd2+. The bulk amount of Cd2+ adsorbed by the metakaolin GP was EDTA extractable, and the adsorbent remained 85% of its adsorption capacity after EDTA desorption for 5 cycles. Luukkonen [32] and Naghsh [58] suggested the efficient metal desorption by 5% NaCl. However, care must be taken since the balancing ions can form a positively charged film on the adsorbent surfaces. El Esweed et al. have achieved ion-exchange based desorption of Cu2+ by 0.1 M NaCl [160]. From all the studies reported, only Cd2+ has been shown to be desorbed at pH > 8 with NaOH solution, achieving 24–84% desorption [64].

Advertisement

5. Environmental impact and costs of treatment with GPs/AAMs

An efficient use of GPs/AAMs in real wastewater treatment practices including economic evaluation is little investigated. Above all, these adsorbents show rather low selectivity, and therefore the ubiquitous metal ions (Na+, Ca2+, Mg2+, Fe3+) present in wastewater solutions demonstrate either competing interaction with the target ions, or the interaction has not been studied [57]. Additionally, for economic and ecological assessment is essential that the adsorbent would be regenerable [168]. To be economically successful, exhausted adsorbents need to pass the non-hazardous leaching criteria of the adsorbed materials, while the amount of waste regenerated should be as little as possible. This means that the adsorption-regeneration cycle needs to be performed as often as possible. And yet, afterwards the adsorbent needs to find end storage place, e.g. in tailing pond, or further use, e.g. as binder, filler, or soil amendment.

Adsorption capacity of a powdered GP is usually higher, but technical implementation of powdered forms requires precise dosing, contact vessel with stirring, solid–liquid separation step, and transfer of exhausted adsorbent to regeneration vessel. The powder can then be regenerated by addition of suitable regenerant, e.g. mild acid, separated, and dried prior to the next adsorption cycle.

Technically, the use of granular forms is an easier option. However, the size of the column vs. wastewater stream can easily become very large, as granules per se, are larger particles and adsorption is a surface process. This puts additional burden on geopolymer production as the overall capacity should be sufficient, and the granules will need to show suitable compressive strength to withstand the gravimetric pressure in the purification column. Conversely, regeneration is technically easily realized by counter flow of regeneration liquid through the column.

Economic evaluation therefore needs to take these considerations into account during CAPEX estimation. OPEX, in turn, is not only the ongoing replacement of exhausted adsorbent, electricity consumed, maintenance, staff, and regeneration chemicals, but also the transportation costs of adsorbents, which can be high at low adsorption capacity.

As a thought experiment, an example of 55 mg/g adsorption capacity of copper adsorbent, with 85% cycling capacity has a 47 mg/g adsorption capacity after desorption cycle, shall be considered. For a mine effluent or process water with 5 mg/L Cu2+ and a flow of 200 m3/h requires about 21 kg adsorbent per hour. The price of GPs is given as 1–1.5 € per kg [103], and as such the treatment costs of merely 1 h would be between 10 and 21 €. Regeneration up to 20 times gives more realistic cost factors, of 0.5–1 € per h, only for adsorbent costs. It becomes quickly clear that without regeneration, high efficiency, and selectivity GPs/AAMs will be too expensive for wastewater treatment.

Advertisement

6. Conclusions

Much work has been done on the adsorption properties of GPs/AAMs towards a wide variety of inorganic pollutants during the last decade. While the effect of competing ions in real water samples remain an issue, the incorporation of new composite materials and the tailoring of reaction conditions have a high potential to increase their selectivity as adsorbents. However, more and more authors have understood the need to regenerate adsorbents and research are being conducted on the recovery of valuable materials, such as metals or nutrients. The recovery of high energy products from side streams utilizing adsorbents made from industrial side streams, will bring circular economy towards the next level. It is also of interest, to cover the costs of water treatment by the revenue of removed materials. While still much work needs to be done, the authors remain confident, that GPs/AAMs will continue to have a prominent place in wastewater treatment.

Advertisement

Acknowledgments

The research was partially funded by European Regional Development Fund (Leverage from the EU, WaterPro project № A74635; Keski-Pohjanmaan liitto/Kainuun liitto/Pohjois-Pohjanmaan liitto) and by Maa- ja vesitekniikan tuki (№ 13-8271-17).

References

  1. 1. Bennich T, Weitz N, Carlsen H. Deciphering the scientific literature on SDG interactions: A review and reading guide. Science of The Total Environment 2020;728:138405. DOI: 10.1016/j.scitotenv.2020.138405
  2. 2. World Bank Group. The Growing Role of Minerals and Metals for a Low Carbon Future. World Bank, Washington, DC; 2017. DOI: 10.1596/28312
  3. 3. Takaluoma E, Samarina T. Valorisation of Sidestream from lithium Refining - Activation of Analcime. Proceedings of the 6th World Congress on New Technologies (NewTech’20), Virtual: AVESTIA; 2020. DOI: 10.11159/icepr20.149
  4. 4. Andrejkovičová S, Sudagar A, Rocha J, Patinha C, Hajjaji W, da Silva EF, et al. The effect of natural zeolite on microstructure, mechanical and heavy metals adsorption properties of metakaolin based geopolymers. Applied Clay Science 2016;126:141–152. DOI: 10.1016/j.clay.2016.03.009
  5. 5. López FJ, Sugita S, Tagaya M, Kobayashi T. Metakaolin-Based Geopolymers for Targeted Adsorbents to Heavy Metal Ion Separation. Journal of Materials Science and Chemical Engineering 2014;2:16–27. DOI: 10.4236/msce.2014.27002
  6. 6. Ariffin N, Abdullah MMAB, Zainol MRRMA, Baltatu MS, Jamaludin L. Effect of Solid to Liquid Ratio on Heavy Metal Removal by Geopolymer-Based Adsorbent. IOP Conf Ser: Mater Sci Eng 2018;374:012045. DOI: 10.1088/1757-899X/374/1/012045
  7. 7. Gollakota ARK, Volli V, Shu C-M. Progressive utilisation prospects of coal fly ash: A review. Sci Total Environ 2019;672:951–989. DOI: 10.1016/j.scitotenv.2019.03.337
  8. 8. Tian Q, Sasaki K. Structural characterizations of fly ash-based geopolymer after adsorption of various metal ions. Environmental Technology 2019;0:1–11. DOI: 10.1080/09593330.2019.1649469
  9. 9. Mužek MN, Svilović S, Ugrina M, Zelić J. Removal of copper and cobalt ions by fly ash-based geopolymer from solutions-equilibrium study. Desalination and Water Treatment 2016;57:10689–10699. DOI: 10.1080/19443994.2015.1040077
  10. 10. Jo M, Soto L, Arocho M, St John J, Hwang S. Optimum mix design of fly ash geopolymer paste and its use in pervious concrete for removal of fecal coliforms and phosphorus in water. Construction and Building Materials 2015;93:1097–1104. DOI: 10.1016/j.conbuildmat.2015.05.034
  11. 11. Novais RM, Buruberri LH, Seabra MP, Labrincha JA. Novel porous fly-ash containing geopolymer monoliths for lead adsorption from wastewaters. J Hazard Mater 2016;318:631–640. DOI: 10.1016/j.jhazmat.2016.07.059
  12. 12. Yao Z. Fly Ash: Sources, Applications and Potential Environmental Impacts. 2014
  13. 13. De Gisi S, Lofrano G, Grassi M, Notarnicola M. Characteristics and adsorption capacities of low-cost sorbents for wastewater treatment: A review. Sustainable Materials and Technologies 2016;9:10–40. DOI: 10.1016/j.susmat.2016.06.002
  14. 14. Ahmaruzzaman M. A review on the utilization of fly ash. Progress in Energy and Combustion Science 2010;36:327–363. DOI: 10.1016/j.pecs.2009.11.003
  15. 15. Ji Z, Pei Y. Geopolymers produced from drinking water treatment residue and bottom ash for the immobilization of heavy metals. Chemosphere 2019;225:579–587. DOI: 10.1016/j.chemosphere.2019.03.056
  16. 16. Ji Z, Pei Y. Bibliographic and visualized analysis of geopolymer research and its application in heavy metal immobilization: A review. Journal of Environmental Management 2019;231:256–267. DOI: 10.1016/j.jenvman.2018.10.041
  17. 17. Maldonado-Alameda À, Giro-Paloma J, Alfocea-Roig A, Formosa J, Chimenos JM. Municipal Solid Waste Incineration Bottom Ash as Sole Precursor in the Alkali-Activated Binder Formulation. Applied Sciences 2020;10:4129. DOI: 10.3390/app10124129
  18. 18. Al-Ghouti MA, Khan M, Nasser MS, Saad KA, Heng OE. Application of geopolymers synthesized from incinerated municipal solid waste ashes for the removal of cationic dye from water. PLOS ONE 2020;15:e0239095. DOI: 10.1371/journal.pone.0239095
  19. 19. Silva RV, de Brito J, Lynn CJ, Dhir RK. Use of municipal solid waste incineration bottom ashes in alkali-activated materials, ceramics and granular applications: A review. Waste Management 2017;68:207–220. DOI: 10.1016/j.wasman.2017.06.043
  20. 20. Chiang YW, Ghyselbrecht K, Santos RM, Meesschaert B, Martens JA. Synthesis of zeolitic-type adsorbent material from municipal solid waste incinerator bottom ash and its application in heavy metal adsorption. Catalysis Today 2012;190:23–30. DOI: 10.1016/j.cattod.2011.11.002
  21. 21. Hong S-H, Shin M-C, Lee J, Lee C-G, Song D-S, Um B-H, et al. Recycling of bottom ash derived from combustion of cattle manure and its adsorption behaviors for Cd(II), Cu(II), Pb(II), and Ni(II). Environ Sci Pollut Res 2020. DOI: 10.1007/s11356-020-11719-7
  22. 22. Zhang HY, Zheng Y, Hu HT, Qi JY. Use of Municipal Solid Waste Incineration Bottom Ash in Adsorption of Heavy Metals. Key Engineering Materials 2011;474–476:1099–1102. DOI: 10.4028/www.scientific.net/KEM.474-476.1099
  23. 23. Joseph A, Snellings R, Van den Heede P, Matthys S, De Belie N. The Use of Municipal Solid Waste Incineration Ash in Various Building Materials: A Belgian Point of View. Materials 2018;11:141. DOI: 10.3390/ma11010141
  24. 24. Zhu W, Rao XH, Liu Y, Yang E-H. Lightweight aerated metakaolin-based geopolymer incorporating municipal solid waste incineration bottom ash as gas-forming agent. Journal of Cleaner Production 2018;177:775–781. DOI: 10.1016/j.jclepro.2017.12.267
  25. 25. Fabricius A-L, Renner M, Voss M, Funk M, Perfoll A, Gehring F, et al. Municipal waste incineration fly ashes: from a multi-element approach to market potential evaluation. Environmental Sciences Europe 2020;32:88. DOI: 10.1186/s12302-020-00365-y
  26. 26. Quina MJ, Bontempi E, Bogush A, Schlumberger S, Weibel G, Braga R, et al. Technologies for the management of MSW incineration ashes from gas cleaning: New perspectives on recovery of secondary raw materials and circular economy. Science of The Total Environment 2018;635:526–542. DOI: 10.1016/j.scitotenv.2018.04.150
  27. 27. Zhuang XY, Chen L, Komarneni S, Zhou CH, Tong DS, Yang HM, et al. Fly ash-based geopolymer: clean production, properties and applications. Journal of Cleaner Production 2016;125:253–267. DOI: 10.1016/j.jclepro.2016.03.019
  28. 28. Wang GC. 11 - Slag use as an aggregate in concrete and cement-based materials. In: Wang GC, editor. The Utilization of Slag in Civil Infrastructure Construction, Woodhead Publishing; 2016, p. 239–74. DOI: 10.1016/B978-0-08-100381-7.00011-2
  29. 29. Siddique R, Iqbal Khan M. Ground Granulated Blast Furnace Slag. In: Siddique R, Khan MI, editors. Supplementary Cementing Materials, Berlin, Heidelberg: Springer; 2011, p. 121–173. DOI: 10.1007/978-3-642-17866-5_3
  30. 30. Gupta VK, Carrott PJM, Carrott MMLR, Suhas. Low-Cost Adsorbents: Growing Approach to Wastewater Treatment—a Review. Critical Reviews in Environmental Science and Technology 2009;39:783–842. DOI: 10.1080/10643380801977610
  31. 31. Medina TJ, Arredondo SP, Corral R, Jacobo A, Zarraga RA, Rosas CA, et al. Microstructure and Pb(2+)Adsorption Properties of Blast Furnace Slag and Fly Ash based Geopolymers. Minerals 2020;10:808. DOI: 10.3390/min10090808
  32. 32. Luukkonen T, Runtti H, Niskanen M, Tolonen E-T, Sarkkinen M, Kemppainen K, et al. Simultaneous removal of Ni(II), As(III), and Sb(III) from spiked mine effluent with metakaolin and blast-furnace-slag geopolymers. Journal of Environmental Management 2016;166:579–588. DOI: 10.1016/j.jenvman.2015.11.007
  33. 33. Tang Q, Wang K, Yaseen M, Tong Z, Cui X. Synthesis of highly efficient porous inorganic polymer microspheres for the adsorptive removal of Pb2+ from wastewater. J Clean Prod 2018;193:351–362. DOI: 10.1016/j.jclepro.2018.05.094
  34. 34. Tang Q, Wang K, Su J, Shen Y, Yang S, Ge Y, et al. Facile fabrication of inorganic polymer microspheres as adsorbents for removing heavy metal ions. Mater Res Bull 2019;113:202–208. DOI: 10.1016/j.materresbull.2019.02.009
  35. 35. Khalid HR, Lee NK, Park SM, Abbas N, Lee HK. Synthesis of geopolymer-supported zeolites via robust one-step method and their adsorption potential. Journal of Hazardous Materials 2018;353:522–533. DOI: 10.1016/j.jhazmat.2018.04.049
  36. 36. Manchisi J, Matinde E, Rowson NA, Simmons MJH, Simate GS, Ndlovu S, et al. Ironmaking and Steelmaking Slags as Sustainable Adsorbents for Industrial Effluents and Wastewater Treatment: A Critical Review of Properties, Performance, Challenges and Opportunities. Sustainability 2020;12:2118. DOI: 10.3390/su12052118
  37. 37. Naidu TS, Sheridan CM, van Dyk LD. Basic oxygen furnace slag: Review of current and potential uses. Minerals Engineering 2020;149:106234. DOI: 10.1016/j.mineng.2020.106234
  38. 38. Nikolić I, Đurović D, Tadić M, Radmilović VV, Radmilović VR. Adsorption kinetics, equilibrium, and thermodynamics of Cu2+ on pristine and alkali activated steel slag. Chemical Engineering Communications 2020;207:1278–1297. DOI: 10.1080/00986445.2019.1685986
  39. 39. Sarkar C, Basu JK, Samanta AN. Removal of Ni2+ ion from waste water by Geopolymeric Adsorbent derived from LD Slag. Journal of Water Process Engineering 2017;17:237–244. DOI: 10.1016/j.jwpe.2017.04.012
  40. 40. Sarkar C, Basu JK, Samanta AN. Synthesis of mesoporous geopolymeric powder from LD slag as superior adsorbent for Zinc (II) removal. Advanced Powder Technology 2018;29:1142–1152. DOI: 10.1016/j.apt.2018.02.005
  41. 41. Sarkar C, Basu JK, Samanta AN. Experimental and kinetic study of fluoride adsorption by Ni and Zn modified LD slag based geopolymer. Chemical Engineering Research and Design 2019;142:165–175. DOI: 10.1016/j.cherd.2018.12.006
  42. 42. Sithole NT, Ntuli F, Okonta F. Fixed bed column studies for decontamination of acidic mineral effluent using porous fly ash-basic oxygen furnace slag based geopolymers. Miner Eng 2020;154:106397. DOI: 10.1016/j.mineng.2020.106397
  43. 43. Sithole NT, Ntuli F, Okonta F. Synthesis and evaluation of basic oxygen furnace slag based geopolymers for removal of metals and sulphates from acidic industrial effluent-column study. J Water Process Eng 2020;37:101518. DOI: 10.1016/j.jwpe.2020.101518
  44. 44. Alhassan SI, He Y, Huang L, Wu B, Yan L, Deng H, et al. A review on fluoride adsorption using modified bauxite: Surface modification and sorption mechanisms perspectives. Journal of Environmental Chemical Engineering 2020;8:104532. DOI: 10.1016/j.jece.2020.104532
  45. 45. Alshaebi F, Yaacob WZ, Samsudin A, Esmail A. Arsenic Adsorption on Bauxite Mineral Using Batch Equilibrium Test. American Journal of Applied Sciences 2009;6. DOI: 10.3844/ajassp.2009.1826.1830
  46. 46. Hertel T, Pontikes Y. Geopolymers, inorganic polymers, alkali-activated materials and hybrid binders from bauxite residue (red mud) – Putting things in perspective. Journal of Cleaner Production 2020;258:120610. DOI: 10.1016/j.jclepro.2020.120610
  47. 47. Norouzi S, Badii K, Doulati Ardejani F. Activated bauxite waste as an adsorbent for removal of Acid Blue 92 from aqueous solutions. Water Sci Technol 2010;62:2491–2500. DOI: 10.2166/wst.2010.514
  48. 48. Taneez M, Hurel C, Marmier N, Lefèvre G. Adsorption of inorganic pollutants on bauxite residues: An example of methodology to simulate adsorption in complex solids mixtures. Applied Geochemistry 2017;78:272–278. DOI: 10.1016/j.apgeochem.2017.01.007
  49. 49. Ren B, Zhao Y, Bai H, Kang S, Zhang T, Song S. Eco-friendly geopolymer prepared from solid wastes: A critical review. Chemosphere 2020:128900. DOI: 10.1016/j.chemosphere.2020.128900
  50. 50. Khalid MK, Agarwal V, Wilson BP, Takaluoma E, Lundstrom M. Applicability of solid process residues as sorbents for the treatment of industrial wastewaters. J Clean Prod 2020;246:118951. DOI: 10.1016/j.jclepro.2019.118951
  51. 51. Demir F, Derun EM. Modelling and optimization of gold mine tailings based geopolymer by using response surface method and its application in Pb2+ removal. J Clean Prod 2019;237:117766. DOI: 10.1016/j.jclepro.2019.117766
  52. 52. Tian Q, Nakama S, Sasaki K. Immobilization of cesium in fly ash-silica fume based geopolymers with different Si/Al molar ratios. Sci Total Environ 2019;687:1127–1137. DOI: 10.1016/j.scitotenv.2019.06.095
  53. 53. Shakouri S, Bayer O, Erdogan ST. Development of silica fume-based geopolymer foams. Constr Build Mater 2020;260:120442. DOI: 10.1016/j.conbuildmat.2020.120442
  54. 54. Duan P, Yan C, Zhou W, Ren D. Development of fly ash and iron ore tailing based porous geopolymer for removal of Cu(II) from wastewater. Ceram Int 2016;42:13507–13518. DOI: 10.1016/j.ceramint.2016.05.143
  55. 55. Panda L, Rath SS, Rao DS, Nayak BB, Das B, Misra PK. Thorough understanding of the kinetics and mechanism of heavy metal adsorption onto a pyrophyllite mine waste based geopolymer. Journal of Molecular Liquids 2018;263:428–441. DOI: 10.1016/j.molliq.2018.05.016
  56. 56. Lemougna PN, Wang K, Tang Q, Melo UC, Cui X. Recent developments on inorganic polymers synthesis and applications. Ceramics International 2016;42:15142–15159. DOI: 10.1016/j.ceramint.2016.07.027
  57. 57. Rasaki SA, Bingxue Z, Guarecuco R, Thomas T, Minghui Y. Geopolymer for use in heavy metals adsorption, and advanced oxidative processes: A critical review. Journal of Cleaner Production 2019;213:42–58. DOI: 10.1016/j.jclepro.2018.12.145
  58. 58. Naghsh M, Shams K. Synthesis of a kaolin-based geopolymer using a novel fusion method and its application in effective water softening. Applied Clay Science 2017;146:238–245. DOI: 10.1016/j.clay.2017.06.008
  59. 59. Samarina T, Takaluoma E. Metakaolin-Based Geopolymers for Removal of Ammonium from Municipal Wastewater, 2019. DOI: 10.11159/icepr19.195
  60. 60. Khalifa AZ, Cizer Ö, Pontikes Y, Heath A, Patureau P, Bernal SA, et al. Advances in alkali-activation of clay minerals. Cement and Concrete Research 2020;132:106050. DOI: 10.1016/j.cemconres.2020.106050
  61. 61. Maleki A, Hajizadeh Z, Sharifi V, Emdadi Z. A green, porous and eco-friendly magnetic geopolymer adsorbent for heavy metals removal from aqueous solutions. J Clean Prod 2019;215:1233–1245. DOI: 10.1016/j.jclepro.2019.01.084
  62. 62. Ghani U, Hussain S, Noor-ul-Amin, Imtiaz M, Ali Khan S. Laterite clay-based geopolymer as a potential adsorbent for the heavy metals removal from aqueous solutions. Journal of Saudi Chemical Society 2020;24:874–884. DOI: 10.1016/j.jscs.2020.09.004
  63. 63. Hayashi S, Onuma T, Kagaya F. Fixing of electrophoretically deposited natural zeolite particles by geopolymer reaction for heavy metal ion adsorbents. Journal of the Ceramic Society of Japan 2017;125:894–898. DOI: 10.2109/jcersj2.17134
  64. 64. Javadian H, Ghorbani F, Tayebi H, Asl SH. Study of the adsorption of Cd (II) from aqueous solution using zeolite-based geopolymer, synthesized from coal fly ash; kinetic, isotherm and thermodynamic studies. Arabian Journal of Chemistry 2015;8:837–849. DOI: 10.1016/j.arabjc.2013.02.018
  65. 65. Wang S, Peng Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chemical Engineering Journal 2010;156:11–24. DOI: 10.1016/j.cej.2009.10.029
  66. 66. Li Y, Li L, Yu J. Applications of Zeolites in Sustainable Chemistry. Chem 2017;3:928–949. DOI: 10.1016/j.chempr.2017.10.009
  67. 67. Delkash M, Ebrazi Bakhshayesh B, Kazemian H. Using zeolitic adsorbents to cleanup special wastewater streams: A review. Microporous and Mesoporous Materials 2015;214:224–241. DOI: 10.1016/j.micromeso.2015.04.039
  68. 68. Yilmaz B, Müller U. Catalytic Applications of Zeolites in Chemical Industry. Top Catal 2009;52:888–895. DOI: 10.1007/s11244-009-9226-0
  69. 69. Montalvo S, Huiliñir C, Borja R, Sánchez E, Herrmann C. Application of zeolites for biological treatment processes of solid wastes and wastewaters – A review. Bioresource Technology 2020;301:122808. DOI: 10.1016/j.biortech.2020.122808
  70. 70. Khaleque A, Alam MM, Hoque M, Mondal S, Haider JB, Xu B, et al. Zeolite synthesis from low-cost materials and environmental applications: A review. Environmental Advances 2020;2:100019. DOI: 10.1016/j.envadv.2020.100019
  71. 71. Rożek P, Król M, Mozgawa W. Geopolymer-zeolite composites: A review. Journal of Cleaner Production 2019;230:557–579. DOI: 10.1016/j.jclepro.2019.05.152
  72. 72. Alshaaer M, Zaharaki D, Komnitsas K. Microstructural characteristics and adsorption potential of a zeolitic tuff–metakaolin geopolymer. Desalination and Water Treatment 2015;56:338–345. DOI: 10.1080/19443994.2014.938306
  73. 73. Deng L, Xu Q, Wu H. Synthesis of Zeolite-like Material by Hydrothermal and Fusion Methods Using Municipal Solid Waste Fly Ash. Procedia Environmental Sciences 2016;31:662–667. DOI: 10.1016/j.proenv.2016.02.122
  74. 74. Visa M. Synthesis and characterization of new zeolite materials obtained from fly ash for heavy metals removal in advanced wastewater treatment. Powder Technology 2016;294:338–347. DOI: 10.1016/j.powtec.2016.02.019
  75. 75. Rios C, Williams C, Castellanos OM. A comparative study on the use of zeolite-like materials in the removal of heavy metals and ammonium from wastewaters. RP 2020;18. DOI: 10.15665/rp.v18i2.2187
  76. 76. Toniolo N, Boccaccini AR. Fly ash-based geopolymers containing added silicate waste. A review. Ceramics International 2017;43:14545–14551. DOI: 10.1016/j.ceramint.2017.07.221
  77. 77. Gasca-Tirado JR, Manzano-Ramírez A, Vazquez-Landaverde PA, Herrera-Díaz EI, Rodríguez-Ugarte ME, Rubio-Ávalos JC, et al. Ion-exchanged geopolymer for photocatalytic degradation of a volatile organic compound. Materials Letters 2014;134:222–224. DOI: 10.1016/j.matlet.2014.07.090
  78. 78. Ge Y, Cui X, Kong Y, Li Z, He Y, Zhou Q. Porous geopolymeric spheres for removal of Cu(II) from aqueous solution: Synthesis and evaluation. Journal of Hazardous Materials 2015;283:244–251. DOI: 10.1016/j.jhazmat.2014.09.038
  79. 79. Zhang M, Zhao M, Zhang G, Nowak P, Coen A, Tao M. Calcium-free geopolymer as a stabilizer for sulfate-rich soils. Applied Clay Science 2015;108:199–207. DOI: 10.1016/j.clay.2015.02.029
  80. 80. Duan J, Li J, Lu Z. One-step facile synthesis of bulk zeolite A through metakaolin-based geopolymer gels. J Porous Mater 2015;22:1519–1526. DOI: 10.1007/s10934-015-0034-6
  81. 81. Tang Q, Ge Y, Wang K, He Y, Cui X. Preparation and characterization of porous metakaolin-based inorganic polymer spheres as an adsorbent. Materials & Design 2015;88:1244–1249. DOI: 10.1016/j.matdes.2015.09.126
  82. 82. Kovářík T, Křenek T, Pola M, Rieger D, Kadlec J, Franče P. Ceramic-like open-celled geopolymer foam as a porous substrate for water treatment catalyst. IOP Conf Ser: Mater Sci Eng 2017;175:012044. DOI: 10.1088/1757-899X/175/1/012044
  83. 83. Ge Y, Cui X, Liao C, Li Z. Facile fabrication of green geopolymer/alginate hybrid spheres for efficient removal of Cu(II) in water: Batch and column studies. Chemical Engineering Journal 2017;311:126–134. DOI: 10.1016/j.cej.2016.11.079
  84. 84. Kara İ, Yilmazer D, Akar ST. Metakaolin based geopolymer as an effective adsorbent for adsorption of zinc(II) and nickel(II) ions from aqueous solutions. Applied Clay Science 2017;139:54–63. DOI: 10.1016/j.clay.2017.01.008
  85. 85. Singhal A, Gangwar BP, Gayathry JM. CTAB modified large surface area nanoporous geopolymer with high adsorption capacity for copper ion removal. Applied Clay Science 2017;150:106–114. DOI: 10.1016/j.clay.2017.09.013
  86. 86. Novais RM, Ascensão G, Tobaldi DM, Seabra MP, Labrincha JA. Biomass fly ash geopolymer monoliths for effective methylene blue removal from wastewaters. Journal of Cleaner Production 2018;171:783–794. DOI: 10.1016/j.jclepro.2017.10.078
  87. 87. Kara I, Tunc D, Sayin F, Akar ST. Study on the performance of metakaolin based geopolymer for Mn(II) and Co(II) removal. Applied Clay Science 2018;161:184–193. DOI: 10.1016/j.clay.2018.04.027
  88. 88. Bravo PI, Malenab RA, Shimizu E, Yu DE, Promentilla MA. Synthesis of geopolymer spheres with photocatalytic activity. MATEC Web Conf 2019;268:04007. DOI: 10.1051/matecconf/201926804007
  89. 89. Novais RM, Carvalheiras J, Tobaldi DM, Seabra MP, Pullar RC, Labrincha JA. Synthesis of porous biomass fly ash-based geopolymer spheres for efficient removal of methylene blue from wastewaters. Journal of Cleaner Production 2019;207:350–362. DOI: 10.1016/j.jclepro.2018.09.265
  90. 90. Yan C, Guo L, Ren D, Duan P. Novel composites based on geopolymer for removal of Pb(II). Materials Letters 2019;239:192–195. DOI: 10.1016/j.matlet.2018.12.105
  91. 91. Samarina T, Takaluoma E. Simultaneous Removal of Nutrients by Geopolymers Made From Industrial By-Products. Proceedings of the 5th World Congress on New Technologies (NewTech’19), Lisbone: AVESTIA; 2019. DOI: 10.11159/ICEPR19.169
  92. 92. El Alouani M, Alehyen S, El Achouri M, Taibi M. Preparation, Characterization, and Application of Metakaolin-Based Geopolymer for Removal of Methylene Blue from Aqueous Solution. Journal of Chemistry 2019;2019:e4212901. DOI: https://doi.org/10.1155/2019/4212901
  93. 93. Yan S, Zhang F, Wang L, Rong Y, He P, Jia D, et al. A green and low-cost hollow gangue microsphere/geopolymer adsorbent for the effective removal of heavy metals from wastewaters. Journal of Environmental Management 2019;246:174–183. DOI: 10.1016/j.jenvman.2019.05.120
  94. 94. Chen X, Niu Z, Zhang H, Guo Y, Liu M, Zhou M. Study on the metakaolin-based geopolymer pervious concrete (MKGPC) and its removal capability of heavy metal ions. International Journal of Pavement Engineering 2019;0:1–12. DOI: 10.1080/10298436.2019.1667499
  95. 95. Bumanis G, Novais RM, Carvalheiras J, Bajare D, Labrincha JA. Metals removal from aqueous solutions by tailored porous waste-based granulated alkali-activated materials. Appl Clay Sci 2019;179:105147. DOI: 10.1016/j.clay.2019.105147
  96. 96. Lan T, Li P, Rehman FU, Li X, Yang W, Guo S. Efficient adsorption of Cd2+ from aqueous solution using metakaolin geopolymers. Environ Sci Pollut Res 2019;26:33555–33567. DOI: 10.1007/s11356-019-06362-w
  97. 97. Chen S, Qi Y, Cossa JJ, Deocleciano Salomao Dos SI. Efficient removal of radioactive iodide anions from simulated wastewater by HDTMA-geopolymer. Progress in Nuclear Energy 2019;117:103112. DOI: 10.1016/j.pnucene.2019.103112
  98. 98. Falah M, MacKenzie KJD, Knibbe R, Page SJ, Hanna JV. New composites of nanoparticle Cu (I) oxide and titania in a novel inorganic polymer (geopolymer) matrix for destruction of dyes and hazardous organic pollutants. Journal of Hazardous Materials 2016;318:772–782. DOI: 10.1016/j.jhazmat.2016.06.016
  99. 99. Yu Z, Song W, Li J, Li Q. Improved simultaneous adsorption of Cu(II) and Cr (VI) of organic modified metakaolin-based geopolymer. Arab J Chem 2020;13:4811–4823. DOI: 10.1016/j.arabjc.2020.01.001
  100. 100. Hua P, Sellaoui L, Franco D, Netto MS, Luiz Dotto G, Bajahzar A, et al. Adsorption of acid green and procion red on a magnetic geopolymer based adsorbent: Experiments, characterization and theoretical treatment. Chemical Engineering Journal 2020;383:123113. DOI: 10.1016/j.cej.2019.123113
  101. 101. Chen H, Zhang YJ, He PY, Li CJ, Li H. Coupling of self-supporting geopolymer membrane with intercepted Cr(III) for dye wastewater treatment by hybrid photocatalysis and membrane separation. Applied Surface Science 2020;515:146024. DOI: 10.1016/j.apsusc.2020.146024
  102. 102. Mladenovic N, Kljajevic L, Nenadovic S, Ivanovic M, Calija B, Gulicovski J, et al. The Applications of New Inorganic Polymer for Adsorption Cadmium from Waste Water. J Inorg Organomet Polym Mater 2020;30:554–563. DOI: 10.1007/s10904-019-01215-y
  103. 103. da Costa Rocha AC, Scaratti G, Moura-Nickel CD, da Silva TL, Gurgel Adeodato Vieira M, Peralta RM, et al. Economical and Technological Aspects of Copper Removal from Water Using a Geopolymer and Natural Zeolite. Water Air Soil Pollut 2020;231:361. DOI: 10.1007/s11270-020-04722-8
  104. 104. Papa E, Mor M, Natali Murri A, Landi E, Medri V. Ice-templated geopolymer beads for dye removal. Journal of Colloid and Interface Science 2020;572:364–373. DOI: 10.1016/j.jcis.2020.03.104
  105. 105. Medri V, Papa E, Mor M, Vaccari A, Natali Murri A, Piotte L, et al. Mechanical strength and cationic dye adsorption ability of metakaolin-based geopolymer spheres. Applied Clay Science 2020;193:105678. DOI: 10.1016/j.clay.2020.105678
  106. 106. Netto MS, Rossatto DL, Jahn SL, Mallmann ES, Dotto GL, Foletto EL. Preparation of a novel magnetic geopolymer/zero–valent iron composite with remarkable adsorption performance towards aqueous Acid Red 97. Chemical Engineering Communications 2020;207:1048–1061. DOI: 10.1080/00986445.2019.1635467
  107. 107. Samarina T, Takaluoma E, Kazmaganbetova A. Geopolymers Supported on Inert Substrate for Phosphate Removal from Natural Waters. Proceedings of the 6th World Congress on New Technologies (NewTech’20), Virtual: AVESTIA; 2020. DOI: 10.11159/icepr20.119
  108. 108. Huang J, Li Z, Zhang J, Zhang Y, Ge Y, Cui X. In-situ synchronous carbonation and self-activation of biochar/geopolymer composite membrane: Enhanced catalyst for oxidative degradation of tetracycline in water. Chemical Engineering Journal 2020;397:125528. DOI: 10.1016/j.cej.2020.125528
  109. 109. Petlitckaia S, Barré Y, Piallat T, Grauby O, Ferry D, Poulesquen A. Functionalized geopolymer foams for cesium removal from liquid nuclear waste. Journal of Cleaner Production 2020;269:122400. DOI: 10.1016/j.jclepro.2020.122400
  110. 110. Lertcumfu N, Jaita P, Thammarong S, Lamkhao S, Tandorn S, Randorn C, et al. Influence of graphene oxide additive on physical, microstructure, adsorption, and photocatalytic properties of calcined kaolinite-based geopolymer ceramic composites. Colloid Surf A-Physicochem Eng Asp 2020;602:125080. DOI: 10.1016/j.colsurfa.2020.125080
  111. 111. Al-Harahsheh MS, Al Zboon K, Al-Makhadmeh L, Hararah M, Mahasneh M. Fly ash based geopolymer for heavy metal removal: A case study on copper removal. Journal of Environmental Chemical Engineering 2015;3:1669–1677. DOI: 10.1016/j.jece.2015.06.005
  112. 112. Meher AK, Das S, Rayalu S, Bansiwal A. Enhanced arsenic removal from drinking water by iron-enriched aluminosilicate adsorbent prepared from fly ash. Desalin Water Treat 2015;57:20944–20956. DOI: 10.1080/19443994.2015.1112311
  113. 113. Zhang K, Zhang D, Zhang K. Arsenic removal from water using a novel amorphous adsorbent developed from coal fly ash. Water Sci Technol 2016;73:1954–1962. DOI: 10.2166/wst.2016.028
  114. 114. Liu Y, Yan C, Zhang Z, Wang H, Zhou S, Zhou W. A comparative study on fly ash, geopolymer and faujasite block for Pb removal from aqueous solution. Fuel 2016;185:181–189. DOI: 10.1016/j.fuel.2016.07.116
  115. 115. Samarina T, Takaluoma E. New Geopolymer Adsorbents for Phosphate Removal from Diluted Solutions and their Applications. Proceedings of the 5th World Congress on New Technologies (NewTech’19), Lisbone: AVESTIA; 2019. DOI: 10.11159/ICEPR19.158
  116. 116. Mužek MN, Svilović S, Zelić J. Kinetic studies of cobalt ion removal from aqueous solutions using fly ash-based geopolymer and zeolite NaX as sorbents. Separation Science and Technology 2016;51:2868–2875. DOI: 10.1080/01496395.2016.1228675
  117. 117. Qiu J, Zhao Y, Xing J, Sun X. Fly ash-based geopolymer as a potential adsorbent for Cr(VI) removal. Desalin Water Treat 2017;70:201–209. DOI: 10.5004/dwt.2017.20493
  118. 118. Lee NK, Khalid HR, Lee HK. Adsorption characteristics of cesium onto mesoporous geopolymers containing nano-crystalline zeolites. Microporous Mesoporous Mat 2017;242:238–244. DOI: 10.1016/j.micromeso.2017.01.030
  119. 119. Maiti M, Sarkar M, Malik MA, Xu S, Li Q, Mandal S. Iron Oxide NPs Facilitated a Smart Building Composite for Heavy-Metal Removal and Dye Degradation. ACS Omega 2018;3:1081–1089. DOI: 10.1021/acsomega.7b01545
  120. 120. Zhang YJ, He PY, Zhang YX, Chen H. A novel electroconductive graphene/fly ash-based geopolymer composite and its photocatalytic performance. Chemical Engineering Journal 2018;334:2459–2466. DOI: 10.1016/j.cej.2017.11.171
  121. 121. Samiullah M, Aslam Z, Rana AG, Abbas A, Ahmad W. Alkali-Activated Boiler Fly Ash for Ni(II) Removal: Characterization and Parametric Study. Water Air Soil Pollut 2018;229:113. DOI: 10.1007/s11270-018-3758-5
  122. 122. Darmayanti L, Notodarmojo S, Damanhuri E, Kadja GTM, Mukti RR. Preparation of alkali-activated fly ash-based geopolymer and their application in the adsorption of copper (II) and zinc (II) ions. MATEC Web Conf 2019;276:06012. DOI: 10.1051/matecconf/201927606012
  123. 123. Onutai S, Kobayashi T, Thavorniti P, Jiemsirilers S. Porous fly ash-based geopolymer composite fiber as an adsorbent for removal of heavy metal ions from wastewater. Materials Letters 2019;236:30–33. DOI: 10.1016/j.matlet.2018.10.035
  124. 124. Soni R, Shukla DP. Synthesis of fly ash based zeolite-reduced graphene oxide composite and its evaluation as an adsorbent for arsenic removal. Chemosphere 2019;219:504–509. DOI: 10.1016/j.chemosphere.2018.11.203
  125. 125. Tian Q, Sasaki K. Application of fly ash-based geopolymer for removal of cesium, strontium and arsenate from aqueous solutions: kinetic, equilibrium and mechanism analysis. Water Sci Technol 2019;79:2116–2125. DOI: 10.2166/wst.2019.209
  126. 126. Siyal AA, Shamsuddin MR, Rabat NE, Zulfiqar M, Man Z, Low A. Fly ash based geopolymer for the adsorption of anionic surfactant from aqueous solution. Journal of Cleaner Production 2019;229:232–243. DOI: 10.1016/j.jclepro.2019.04.384
  127. 127. Darmayanti L, Kadja GTM, Notodarmojo S, Damanhuri E, Mukti RR. Structural alteration within fly ash-based geopolymers governing the adsorption of Cu2+ from aqueous environment: Effect of alkali activation. J Hazard Mater 2019;377:305–314. DOI: 10.1016/j.jhazmat.2019.05.086
  128. 128. Shao N, Tang S, Li S, Chen H, Zhang Z. Defective analcime/geopolymer composite membrane derived from fly ash for ultrafast and highly efficient filtration of organic pollutants. Journal of Hazardous Materials 2020;388:121736. DOI: 10.1016/j.jhazmat.2019.121736
  129. 129. Abdelrahman EA, Abou El-Reash YG, Youssef HM, Kotp YH, Hegazey RM. Utilization of rice husk and waste aluminum cans for the synthesis of some nanosized zeolite, zeolite/zeolite, and geopolymer/zeolite products for the efficient removal of Co(II), Cu(II), and Zn(II) ions from aqueous media. Journal of Hazardous Materials 2021;401:123813. DOI: 10.1016/j.jhazmat.2020.123813
  130. 130. Mondal SK, Welz A, Rezaei F, Kumar A, Okoronkwo MU. Structure-Property Relationship of Geopolymers for Aqueous Pb Removal. ACS Omega 2020;5:21689–21699. DOI: 10.1021/acsomega.0c02591
  131. 131. Novais RM, Carvalheiras J, Seabra MP, Pullar RC, Labrincha JA. Highly efficient lead extraction from aqueous solutions using inorganic polymer foams derived from biomass fly ash and metakaolin. J Environ Manage 2020;272:111049. DOI: 10.1016/j.jenvman.2020.111049
  132. 132. Zhang YJ, He PY, Yang MY, Kang L. A new graphene bottom ash geopolymeric composite for photocatalytic H2 production and degradation of dyeing wastewater. International Journal of Hydrogen Energy 2017;42:20589–20598. DOI: 10.1016/j.ijhydene.2017.06.156
  133. 133. Zhang YJ, Yang MY, Zhang L, Zhang K, Kang L. A new graphene/geopolymer nanocomposite for degradation of dye wastewater. Integrated Ferroelectrics 2016;171:38–45. DOI: 10.1080/10584587.2016.1171178
  134. 134. Runtti H, Luukkonen T, Niskanen M, Tuomikoski S, Kangas T, Tynjälä P, et al. Sulphate removal over barium-modified blast-furnace-slag geopolymer. Journal of Hazardous Materials 2016;317:373–384. DOI: 10.1016/j.jhazmat.2016.06.001
  135. 135. He PY, Zhang YJ, Chen H, Liu LC. Development of an eco-efficient CaMoO4/electroconductive geopolymer composite for recycling silicomanganese slag and degradation of dye wastewater. Journal of Cleaner Production 2019;208:1476–1487. DOI: 10.1016/j.jclepro.2018.10.176
  136. 136. Chen F, Wang K, Shao L, Muhammad Y, Wei Y, Gao F, et al. Synthesis of Fe2O3-modified porous geopolymer microspheres for highly selective adsorption and solidification of F− from waste-water. Composites Part B: Engineering 2019;178:107497. DOI: 10.1016/j.compositesb.2019.107497
  137. 137. Liu Z, Li L, Shao N, Hu T, Han L, Wang D. Geopolymerization enhanced hydrothermal synthesis of analcime from steel slag and CFBC fly ash and heavy metal adsorption on analcime. Environ Technol 2020;41:1753–1765. DOI: 10.1080/09593330.2018.1545805
  138. 138. Wang K, Lei H, Muhammad Y, Chen F, Gao F, Wei Y, et al. Controlled preparation of cerium oxide loaded slag-based geopolymer microspheres (CeO2@SGMs) for the adsorptive removal and solidification of F− from acidic waste-water. Journal of Hazardous Materials 2020;400:123199. DOI: 10.1016/j.jhazmat.2020.123199
  139. 139. Yan S, He P, Jia D, Wang Q, Liu J, Yang J, et al. Synthesis of novel low-cost porous gangue microsphere/geopolymer composites and their adsorption properties for dyes. International Journal of Applied Ceramic Technology 2018;15:1602–14. DOI: https://doi.org/10.1111/ijac.13045
  140. 140. Al-Zboon KK, Al-smadi BM, Al-Khawaldh S. Natural Volcanic Tuff-Based Geopolymer for Zn Removal: Adsorption Isotherm, Kinetic, and Thermodynamic Study. Water Air Soil Pollut 2016;227:248. DOI: 10.1007/s11270-016-2937-5
  141. 141. Minelli M, Medri V, Papa E, Miccio F, Landi E, Doghieri F. Geopolymers as solid adsorbent for CO2 capture. Chemical Engineering Science 2016;148:267–274. DOI: 10.1016/j.ces.2016.04.013
  142. 142. Li C, He Y, Tang Q, Wang K, Cui X. Study of the preparation of CdS on the surface of geopolymer spheres and photocatalyst performance. Materials Chemistry and Physics 2016;178:204–210. DOI: 10.1016/j.matchemphys.2016.05.013
  143. 143. Abdelrahman EA, Subaihi A. Application of Geopolymers Modified with Chitosan as Novel Composites for Efficient Removal of Hg(II), Cd(II), and Pb(II) Ions from Aqueous Media. J Inorg Organomet Polym Mater 2020;30:2440–2463. DOI: 10.1007/s10904-019-01380-0
  144. 144. Demir F, Moroydor Derun E. Usage of gold mine tailings based geopolymer on Cu 2+ adsorption from water. Main Group Chemistry 2019;18:467–476. DOI: 10.3233/MGC-190801
  145. 145. Wang J, Zhao Y, Zhang P, Yang L, Xu H, Xi G. Adsorption characteristics of a novel ceramsite for heavy metal removal from stormwater runoff. Chinese Journal of Chemical Engineering 2018;26:96–103. DOI: 10.1016/j.cjche.2017.04.011
  146. 146. Paranaque JEL, Maguyon-Detras MC, Migo VP, Alfafara CG. Chromium removal from chrome-tannery effluent after alkaline precipitation by adsorption using municipal solid waste-derived activated biochar. 26th Regional Symposium on Chemical Engineering (rsce 2019), vol. 778, Bristol: Iop Publishing Ltd; 2020, p. 012134. DOI: 10.1088/1757-899X/778/1/012134
  147. 147. Panda L, Jena SK, Rath SS, Misra PK. Heavy metal removal from water by adsorption using a low-cost geopolymer. Environ Sci Pollut Res 2020;27:24284–24298. DOI: 10.1007/s11356-020-08482-0
  148. 148. Bakharev T. Thermal behaviour of geopolymers prepared using class F fly ash and elevated temperature curing. Cement and Concrete Research 2006;36:1134–1147. DOI: 10.1016/j.cemconres.2006.03.022
  149. 149. Luukkonen T, Tolonen E-T, Runtti H, Kemppainen K, Perämäki P, Rämö J, et al. Optimization of the metakaolin geopolymer preparation for maximized ammonium adsorption capacity. J Mater Sci 2017;52:9363–9376. DOI: 10.1007/s10853-017-1156-9
  150. 150. Siyal AA, Shamsuddin MR, Khan MI, Rabat NE, Zulfiqar M, Man Z, et al. A review on geopolymers as emerging materials for the adsorption of heavy metals and dyes. Journal of Environmental Management 2018;224:327–339. DOI: 10.1016/j.jenvman.2018.07.046
  151. 151. Luukkonen T, Heponiemi A, Runtti H, Pesonen J, Yliniemi J, Lassi U. Application of alkali-activated materials for water and wastewater treatment: a review. Rev Environ Sci Biotechnol 2019;18:271–297. DOI: 10.1007/s11157-019-09494-0
  152. 152. Luukkonen T, Věžníková K, Tolonen E-T, Runtti H, Yliniemi J, Hu T, et al. Removal of ammonium from municipal wastewater with powdered and granulated metakaolin geopolymer. Environmental Technology 2018;39:414–423. DOI: 10.1080/09593330.2017.1301572
  153. 153. Tan TH, Mo KH, Ling T-C, Lai SH. Current development of geopolymer as alternative adsorbent for heavy metal removal. Environmental Technology & Innovation 2020;18:100684. DOI: 10.1016/j.eti.2020.100684
  154. 154. Ducman V, Korat L. Characterization of geopolymer fly-ash based foams obtained with the addition of Al powder or H2O2 as foaming agents. Materials Characterization 2016;113:207–213. DOI: 10.1016/j.matchar.2016.01.019
  155. 155. Landi E, Medri V, Papa E, Dedecek J, Klein P, Benito P, et al. Alkali-bonded ceramics with hierarchical tailored porosity. Applied Clay Science 2013;73:56–64. DOI: 10.1016/j.clay.2012.09.027
  156. 156. Li Q, He Y, Xu M, Liu J, He M, Huang L, et al. Study on the removal of Ca2+ and Mg2+ in water by the geopolymer-based inorganic membrane 2017;48:01215–20. DOI: 10.3969/j.issn.1001-9731.2017.01.039
  157. 157. Ge Y, Yuan Y, Wang K, He Y, Cui X. Preparation of geopolymer-based inorganic membrane for removing Ni2+ from wastewater. Journal of Hazardous Materials 2015;299:711–718. DOI: 10.1016/j.jhazmat.2015.08.006
  158. 158. Davidovits J. 30 years of successes and failures in geopolymer applications. Market trends and potential breakthroughs. Geopolymer 2002 Conference 2002:1–16
  159. 159. Cheng TW, Lee ML, Ko MS, Ueng TH, Yang SF. The heavy metal adsorption characteristics on metakaolin-based geopolymer. Applied Clay Science 2012;56:90–96. DOI: 10.1016/j.clay.2011.11.027
  160. 160. El-Eswed B, Alshaaer M, Yousef RI, Hamadneh I, Khalili F. Adsorption of Cu(II), Ni(II), Zn(II), Cd(II) and Pb(II) onto Kaolin/Zeolite Based- Geopolymers. Advances in Materials Physics and Chemistry 2013;2:119–125. DOI: 10.4236/ampc.2012.24B032
  161. 161. Luukkonen T, Sarkkinen M, Kemppainen K, Rämö J, Lassi U. Metakaolin geopolymer characterization and application for ammonium removal from model solutions and landfill leachate. Applied Clay Science 2016;119:266–276. DOI: 10.1016/j.clay.2015.10.027
  162. 162. Bai C, Colombo P. Processing, properties and applications of highly porous geopolymers: A review. Ceramics International 2018;44:16103–16118. DOI: 10.1016/j.ceramint.2018.05.219
  163. 163. Bai C, Franchin G, Elsayed H, Zaggia A, Conte L, Li H, et al. High-porosity geopolymer foams with tailored porosity for thermal insulation and wastewater treatment. Journal of Materials Research 2017;32:3251–3259. DOI: 10.1557/jmr.2017.127
  164. 164. Huang X, Zhao H, Hu X, Liu F, Wang L, Zhao X, et al. Optimization of preparation technology for modified coal fly ash and its adsorption properties for Cd2+. J Hazard Mater 2020;392:122461. DOI: 10.1016/j.jhazmat.2020.122461
  165. 165. Zhang YJ, Liu LC, Xu Y, Wang YC, Xu DL. A new alkali-activated steel slag-based cementitious material for photocatalytic degradation of organic pollutant from waste water. Journal of Hazardous Materials 2012;209–210:146–150. DOI: 10.1016/j.jhazmat.2012.01.001
  166. 166. Samarina T, Takaluoma E. Ammonia Recovery from Wastewaters by Adsorption with subsequent Transmembrane Chemical Absorption. Proceedings of the 6th World Congress on New Technologies (NewTech’20), Virtual: AVESTIA; 2020. DOI: 10.11159/icepr20.146
  167. 167. Takaluoma E, Samarina T, Mwandila G, Kabondo Leonard L, Nyirenda Kawunga K, Mwaanga P. Selective Recovery of Copper and Cobalt from Mine Effluent. IMWA 2020 Mine water solutions, New Zealand: IMWA; 2020, p. 181186
  168. 168. Lata S, Singh PK, Samadder SR. Regeneration of adsorbents and recovery of heavy metals: a review. Int J Environ Sci Technol 2015;12:1461–1478. DOI: 10.1007/s13762-014-0714-9

Written By

Tatiana Samarina, Esther Takaluoma and Outi Laatikainen

Submitted: 13 October 2020 Reviewed: 10 March 2021 Published: 28 March 2021