Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Journal of Cleaner Production 230 (2019) 557e579

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Review

Geopolymer-zeolite composites: A review


_ *, Magdalena Kro
Piotr Rozek  l, Włodzimierz Mozgawa
AGH University of Science and Technology, Faculty of Materials Science and Ceramics, Al. Mickiewicza 30, 30-059, Krakow, Poland

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

Article history: Although the formation of zeolites in geopolymers have been mentioned in many works, a particular
Received 15 January 2019 interest in geopolymer-zeolite composites started to increase very recently. Such hybrid materials con-
Received in revised form nect the advantageous properties of both constituents: geopolymer serves as a strong and durable
13 May 2019
support for zeolites, when zeolites provide high surface area, porosity and adsorption capacity. The
Accepted 14 May 2019
Available online 16 May 2019
composites can be therefore used as bulk-type adsorbents as well as membranes in separation and
pervaporation processes. The possibility of the utilization of industrial wastes for their production is also
a sustainability benefit. This review article highlights advances in the field of geopolymer-zeolite com-
Keywords:
Geopolymer
posites manufacturing and applications. The resistance of zeolites in geopolymer matrices, methods of
Alkali-activation their detection and effect of their presence on the properties of hybrid materials are discussed as well.
Zeolite © 2019 Elsevier Ltd. All rights reserved.
Geopolymer-zeolite
Self-supporting
Membrane

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
1.1. Geopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
1.2. Zeolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
1.3. Geopolymer-zeolite composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
2. Synthesis of geopolymer-zeolite composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
2.1. Zeolites as raw materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
2.2. Direct zeolites formation in the matrix of geopolymer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561
2.3. Hydrothermal treatment of geopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
2.4. Mechanism of zeolites formation in geopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
3. Resistance of zeolites in the geopolymeric matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
4. Methods used to detect zeolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
5. Applications of geopolymer-zeolite composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
6. Influence of zeolites on the properties of geopolymer-zeolite composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
7. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575
Declarations of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575

1. Introduction

Geopolymers are generally understood as alkali-activated alu-


minosilicates. They may be considered as an inorganic two-
component system which consists of: (1) a reactive solid source
* Corresponding author. of SiO2 and Al2O3, and (2) an alkaline activation solution (Buchwald
_
E-mail address: prozek@agh.edu.pl (P. Rozek).

https://doi.org/10.1016/j.jclepro.2019.05.152
0959-6526/© 2019 Elsevier Ltd. All rights reserved.
558 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

to fully amorphous (Provis et al., 2005). It exhibits high chemical


List of abbreviations: and fire resistance as well as relatively high mechanical strength
(Zhang et al., 2018). Therefore, geopolymers can be used as con-
BA bottom ash struction and building materials, with a prospect of being an
BET Brunauer-Emmett-Teller environmentally-friendly Portland cement replacement.
C-A-S-H calcium aluminosilicate hydrate Geopolymer-stabilized aggregate bases, in turn, can be used as
CFB circulating fluidized bed stabilizing agents in flexible pavement applications (Hu et al.,
FA fly ash 2018b). Another important application of geopolymers is the
FS fumed silica adsorption of heavy metals and dyes, which is the result of their
FT-IR Fourier-transform infrared spectroscopy porous structure and the presence of negative charges, located on
BFS blast furnace slag aluminum tetrahedra, as well as their mechanical stability, cost-
HT hydrothermal treatment effectiveness, eco-friendliness and high efficiency (Siyal et al.,
MK metakaolin 2018).
MAS NMR magic angle spinning nuclear magnetic resonance The product of geopolymerization is amorphous N-A-S-H gel
N-A-S-H sodium aluminosilicate hydrate (sodium aluminosilicate hydrate). In the initial stage of the acti-
RHA rice hull (husk) ash vation it forms in as Al-rich gel, gradually becoming richer in Si
RT room temperature with the reaction progress, when more SieO groups dissolve and
SEM scanning electron microscope the concentration of silicon in the reaction medium rises (Granizo
SA sodium aluminate et al., 2014). The 3D network structure of geopolymers consists of
SH sodium hydroxide SieOeAl framework (spatially connected chains of [SiO4] and
WG water glass (sodium silicate solution) [AlO4] tetrahedra, sharing oxygen corners) and charge-balancing
XRD X-ray diffraction metal cations. When a raw material with some content of CaO is
used for the synthesis of geopolymers, two types of gels coexist,
i.e. N-A-S-H and CeSeH (calcium silicate hydrate) or its Al-
substituted C-A-S-H form (Khalid et al., 2018). It was also sug-
et al., 2011). The mechanism responsible for geopolymer formation gested that at low alkalinities both N-A-S-H and CeSeH gels form,
e geopolymerization e involves the dissolution of silicon and while at high alkalinities N-A-S-H is a predominant phase with
aluminum atoms from the source material, reorientation of pre- some precipitates of calcium compounds within (Zhang et al.,
cursor ions, and their condensation to form a final solid product. In 2010). The mechanism of geopolymerization involves three steps
addition to an amorphous gel, geopolymers may contain some (Krol et al., 2018): (1) dissolution of aluminosilicate raw material;
amounts of unreacted material as well as newly formed crystalline (2) decrease in the content of polymeric aluminosilicate species
phases, especially zeolites. Such products can be more properly with increase in orthosilicate phases; (3) condensation of tetra-
termed as geopolymer-zeolite composites. These hybrid materials hedra and disappearance of polarized forms. Depending on the
connect the properties of both constituents, exhibiting synergistic alkali-activator used in the synthesis, the speed of the first step
effect. increases in order Li < Na < K (Kro l et al., 2018). The geopolymeric
gel is often considered as zeolite-like phase, zeolite precursors, or
1.1. Geopolymers metastable amorphous zeolites, unable to crystallize due to un-
favorable conditions (Sturm et al., 2016). The Ostwald Law of
In the synthesis of geopolymers, two most commonly used Successive Transformations states that a phase does not always
aluminosilicate precursors are metakaolin and coal fly ash. Meta- transform directly into the most stable state, but often changes
kaolin is an anhydrous aluminosilicate obtained by the thermal into a metastable state of greater likeness to itself (Rees et al.,
decomposition of kaolin, a naturally occurring clay containing 2008).
kaolinite mineral [Al2Si2O5(OH)4] and trace amounts of quartz and
other minerals. Hydroxyl ions are strongly bonded to the alumi- 1.2. Zeolites
nosilicate framework of kaolinite and can be eliminated at tem-
peratures above 550  C; dehydroxylation causes a considerable Zeolites are crystalline, hydrated tectoaluminosilicates with a
atomic rearrangement (Ferna ndez-Jimenez et al., 2008). The utili- specific framework structure, consisting of silica and alumina
zation of industrial by-products and wastes for the synthesis of tetrahedra linked by shared oxygen atoms, and containing well-
materials that can be applied for different purposes is particularly defined channels and chambers, filled with ions and water mole-
important. For instance, the global production of coal combustion cules. Such structure makes their physical and chemical properties
products, including fly ash, was 780 million tonnes in 2010, and its unique, which results with a very wide range of their practical
utilization rate was around 50% (Heidrich et al., 2013). Therefore, applications (Krol et al., 2012). Zeolites are based on TO4 tetrahedra
the preparation of coal fly ash-based geopolymers provides a sus- (T is an aluminum or silicon atom), which can be regarded as pri-
tainable way for the disposal of this by-product, which remains a mary building units (PBU). Linkages between larger number of
worldwide challenge. Another industrial by-products used for tetrahedra are called secondary building units (SBU) (Baerlocher
geopolymer synthesis are, for instance, red mud (Hu et al., 2018a; and McCusker, 2018).
Nie et al., 2019), slag (Sun and Vollpracht, 2018), and biomass fly As mentioned above, zeolites can be obtained in the matrix of
ash (De Rossi et al., 2018). An aluminosilicate precursor has to be geopolymer as a direct product, forming parallel to geo-
alkali-activated. For this purpose alkaline hydroxides, silicates, polymerization. Such an approach manifests sustainability benefits
aluminates, carbonates, sulphates, or their combinations are used since the synthesis of geopolymers is more energy- and time-
(Buchwald et al., 2011). Another important factor for the activation efficient compared to the conventional zeolite synthesis, which in
is heat providing, due to the activation barrier that has to be sur- most cases requires autoclave hydrothermal reactor. Moreover,
passed for the reaction to take place (Bakharev, 2005a). A final solid standard hydrothermal synthesis of zeolites from fly ash fail to
product possesses regions varying in degrees of crystallinity, produce zeolites with the desired structural crystallinity (Alzeer
ranging from highly crystalline, nanocrystalline, or polycrystalline and MacKenzie, 2018). The differences in the synthesis of
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 559

geopolymers and zeolites are also in reaction mixtures, i.e. the forming the sodalite framework. Other non-framework compo-
liquid/solid ratios and alkalinities used. Zeolites generally form in nents, such as Naþ, OH, and H2O, are contained in the b-cages of
systems with H2O/SiO2 molar ratios of 10e100 and OH/SiO2 of the structure (Oh et al., 2011a). Although hydroxysodalite is a
2e20, in comparison to geopolymers, which employ H2O/SiO2 of feldspathoid, it is closely related to zeolites. In this review
2e10 and OH/SiO2 of 0.1e0.5 (Rees et al., 2008). Table 1 presents hydroxysodalite refers to sodalite hydrate.
the structures codes, SBUs and chemical formulas of zeolites Hydroxycancrinite, Na8[AlSiO4]6(OH)2$2H2O, represents the
commonly appearing in geopolymeric matrices. cancrinite framework in ABC-6 family (Oh et al., 2011b). 19 mem-
As shown in Fig. 1 hydroxysodalite, faujasite, and zeolite A are bers of zeolite framework types can be described as ABC-6 family.
built of sodalite cages (SOD-cage) formed by connected single four- Other zeolites belonging to the ABC-6 family are sodalite, chabazite,
and six-membered rings (SBU: S4R, S6R) (Buchwald et al., 2011). levyne, and offretite. This family is characterized by a basic struc-
Zeolite A (LTA framework) possesses the so-called a-cage tural layer consisting of six-membered rings of Si- and Al-
comprised of 8 cuboctahedra linked by 12 cuboids. Due to its tetrahedra. The difference between zeolites representing the ABC-
microstructure, it has excellent ion-exchange capacities, satisfac- 6 family is in the stacking sequence of this layer. If the position of
tory hydrophilic properties, and is environmentally neutral, so it is one six-membered ring on one layer is called ‘A’ and the two
widely used in household detergents and ion-exchange mem- possible positions of the next layer above ‘A’ are called ‘B’ and ‘C’, all
branes (Duan et al., 2015; Yan et al., 2012). In LTA, each SOD-cage is the stacking sequences of the basic layers can be described as ‘AB’,
connected with six nearest neighboring SOD-cages by double T4- ‘ABC’ or ‘AABB’, etc. Hydroxycancrinite is in ‘ABAB’ arrangement
rings (D4R) (Buchwald et al., 2011). and chabazite in ‘AABBCC’ arrangement (Oh et al., 2010).
Faujasite-structure zeolites (zeolite X, Y, and natural faujasite)
are also interesting zeolites with their three-dimensional pores
1.3. Geopolymer-zeolite composites
structure, consisting of sodalite cages, which are connected into
double six-membered rings (D6R). Faujasite shows an open
Geopolymer-zeolite composites may merge interesting proper-
framework and high cation exchange capacity. Thus, faujasite has
ties of both aluminosilicate materials. While geopolymeric gel
many different applications in the adsorption of heavy metals,
serves as a strong and durable support for zeolites, zeolites provide
separation of small gas or liquid molecules, and as a catalyst in fluid
higher surface area, porosity and cation exchange capacity (Provis
catalysis cracking (Liu et al., 2016a). In faujasite structures, a SOD-
et al., 2005). Such zeolite-geopolymer bulk materials provide
cage is connected with four nearest neighboring SOD-cages by
inter-connected and multiscale distributed pores as a combination
double T6-rings (D6R) (Buchwald et al., 2011).
of zeolite microporosity and geopolymer meso- and macropores
In sodalite each SOD-cage is connected with six nearest neigh-
(Takeda et al., 2013; Papa et al., 2018). A synergistic effect of zeolite
boring SOD-cages by common T4-rings (Buchwald et al., 2011).
crystals formation in the matrix of geopolymer, which serves as a
Sodalite b-cage frame is indicated by [AlSiO4]66, a basic unit
strong support, results in the potential for applications, e.g. as

Table 1
Zeolite structures commonly appearing in geopolymeric matrices (Baerlocher and McCusker, 2018; Mozgawa, 2001).

Zeolite structure Zeolite SBU Chemical formula (sodium form)

ANA analcime S4R Na[AlSi2O6]$H2O


CAN hydroxycancrinite S6R Na8[AlSiO4]6(OH)2$2H2O
CHA chabazite (herschelite) D6R Na[AlSi2O6]$3H2O
FAU zeolite X D6R Na2[Al2Si2.4O8.8]$6.7H2O
zeolite Y D6R Na1.88[Al2Si4.8O13.54]$9H2O
GIS zeolite NaeP1 S4R Na3.6[Al3.6Si12.4O32]$12H2O
LTA zeolite A D4R Na2[Al2Si1.85O7.7]$5H2O
NAT natrolite 4e1 Na2[Al2Si3O8]$2H2O
SOD hydroxysodalite (sodalite hydrate) S6R Na6[AlSiO4]6$8H2O

Fig. 1. Structure of zeolites.


560 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

monolithic membranes (De Rossi et al., 2018). Studies on the syn- review the manufacturing approaches of these hybrid materials are
thesis of these hybrid materials are still limited. presented (Section 2), the resistance of zeolites in composites
A significant hydrolysis of amorphous phase is necessary to form (Section 3) is discussed as well as their characterization methods
zeolitic crystals from raw materials, which usually requires an (Section 4), the applications (Section 5) and the influence of zeolites
alkaline reaction system, high solution to solid ratio, and elevated on the properties of hybrid materials (Section 6).
temperature (above 50 ) (Van Deventer et al., 2006). Although the
alkaline activation of metakaolin and fly ash results in aluminosil- 2. Synthesis of geopolymer-zeolite composites
icate gel as a main reaction product, differences in the composition
and microstructure of a raw material affect the microstructure of In this section, a comprehensive review of geopolymer-zeolite
the final product. The utilization of metakaolin results in a very composites manufacturing is presented, including three main ap-
homogeneous matrix, a low Si/Al ratio of a gel, and a high degree of proaches made for this purpose: (1) using zeolites as geo-
zeolitization. Fly ash, in turn, induces the formation of heteroge- polymerization precursors, (2) zeolites formation as a result of the
neous matrices with a higher content of unreacted material, a reaction parallel to geopolymerization, and (3) hydrothermal
higher Si/Al ratio of a gel, and a smaller percentage of zeolites treatment of geopolymers. The role of several factors is discussed,
(Ferna ndez-Jime nez et al., 2008). for instance raw materials, activator, additives, and curing regime.
Provis et al. (2005) reexamined the results concerning the na-
ture of geopolymer gel. This phase should be considered as nano- 2.1. Zeolites as raw materials
particulate with nanometer-scale crystal particles (5 nm in
diameter) surrounded by a secondary continuous phase and/or A simple way to obtain zeolitic phases in geopolymers is the
regions of nanoporosity. Oh et al., 2010, 2011a, 2011b stated that the utilization of zeolites as raw materials. It is possible due to their
product of geopolymerization (N-A-S-H gel) may be regarded as a aluminosilicate nature. Such an approach has been applied with
zeolite precursor, namely a disordered form of the ABC-6 family both natural and synthetic zeolites, using them not only as addi-
group of zeolites (or a single layer of the ABC-6 family structure, tives or fillers with conventional materials such as fly ash or met-
Fig. 2), with hydroxysodalite and hydroxycancrinite as the best akaolin, but also as geopolymerization precursors, contributing to
candidates representing this type of a structure due to their the formation of N-A-S-H gel. In some studies, however, zeolitic
simplest atomic arrangement (AB). The similarity level is as com- phases dissolved completely during the process of alkali-activation,
parable as tobermorite and jennite are to C-(A)-S-H in the cement which resulted in the absence of diffraction peaks corresponding to
science. The amorphous precursor gel from which zeolites crys- zeolites in the diffraction patterns of final materials (Auqui et al.,
tallizes contains many quasi-crystalline regions, acting as sites for 2017; Lynch et al., 2018; Hu et al., 2009; Yousef et al., 2009, 2012;
zeolite nucleation, and their growth begins with the formation of Ulloa et al., 2018). Therefore, the utilization of zeolite as a raw
nanosized crystallites within the amorphous gel particles (Provis material does not guarantee that a final product will be a
et al., 2005). geopolymer-zeolite composite.
Although the formation of zeolites in geopolymers have been Since zeolites utilized for the synthesis of geopolymers usually
mentioned in many papers, there are only several papers that dissolve to a greater or lesser extent, the intensities of diffraction
directly deal with the production, characterization, application and peaks corresponding to zeolites decrease after alkali-activation,
development of geopolymer-zeolite composites (De Rossi et al., which was observed, e.g. in the case of Mexican zeolite (mainly
2018; Khalid et al., 2018; Sturm et al., 2016, 2018; Papa et al., clinoptilolite with minor heulandite part) (Villa et al., 2010), Jor-
2018; Lee et al., 2016a, 2017; Greiser et al., 2017; Minelli et al., danian zeolitic tuff (mainly phillipsite) (El-Eswed et al., 2015;
2018). Therefore, the papers only mentioning the appearance of Alshaaer et al., 2016), and Ecuadorian zeolite-rich tuff (mordenite
zeolites in geopolymers are also referred to in this article. In this e 70%) (Baykara et al., 2017). The degree of zeolite dissolution can
vary significantly, e.g. the amount of mordenite in the study
(Baykara et al., 2017) varied from 30 to 60 wt% depending on acti-
vator concentration. The type of activator also affects the degree of
zeolite dissolution during geopolymerization. Nikolov et al. (2017)
used clinoptilolite-rich (75%) natural zeolite as aluminosilicate
precursor. Zeolite phase was detected in the obtained materials
when sodium carbonate and water glass were used as activator,
while the activation with sodium hydroxide caused the disap-
pearance of clinoptilolite peaks. The compressive strength was
measured for mortars and was relatively low (1e4 MPa) due to the
fact that geopolymers were cured at ambient temperature. Never-
theless, the authors stated that natural zeolite-based geopolymers
can be used as coatings and plasters due to relatively high adhesion
strength (over 2 MPa).
It can be stated that the part of zeolite transforms into geo-
polymeric gel and contributes to the development of compressive
strength (Villa et al., 2010), while unreacted zeolitic phase is still
present in a newly-formed geopolymeric matrix and enhances
other properties, such as heavy metals immobilization capacity, like
in the study (El-Eswed et al., 2015): heavy metals cations could
replace exchangeable sodium cations in the framework of zeolite
phillipsite. The positive effect of natural zeolite on the compressive
Fig. 2. Schematic drawing of basic layer of the ABC-6 family consisting of 6-membered
strength was also observed in the case of geopolymers on the basis
rings (dark gray indicates Al-tetrahedra; light gray indicates Si-based tetrahedra). of metakaolin with the addition of clinoptilolite as a filler
Reproduced with permission from Ref. (Oh et al., 2010). Copyright 2010 Elsevier. (Andrejkovi  et al., 2016). The diffraction peaks corresponding
cova
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 561

to clinoptilolite were registered for the geopolymers regardless the geopolymers. The diffraction peaks corresponding to zeolite A
content of zeolite (25, 50 and 70 wt%). The similar effect was were recorded for the prepared materials, so this zeolite was pre-
observed when additionally cork waste residue was added (20 wt%) sent in geopolymer matrix in unreacted form, positively affecting
to metakaolin-zeolite geopolymers (Sudagar et al., 2018). formaldehyde adsorption capacity.
Synthetic zeolites are also subjected to alkali-activation or uti- Zeolitic wastes were used for the synthesis of geopolymers as
lized as additives to conventional metakaolin- or fly ash-based well. Rodríguez et al. (2013a) investigated a spent fluid catalytic
geopolymers. Papa et al. (2018) used commercially available pow- cracking catalyst as the precursor of geopolymerization. It con-
der of zeolite Na13X as a filler for metakaolin-based geopolymers in tained quartz, mullite as well as zeolitic phases: mordenite, zeolite
order to merge the microporosity of zeolite with the mesoporosity ZSM-5 and dehydrated dealuminated zeolite US-Y (FAU type). As a
of the geopolymer matrix. Prior to mixing, zeolite form was result of alkali-activation, zeolite US-Y dissolved regardless alka-
changed from dehydrated to hydrated to avoid the influence on linity, while high alkalinity caused also the dissolution of mor-
geopolymerization process. The hydration of zeolite in geopolymer denite. A high dissolution degree (reactivity) of zeolite Y is related
slurry could affect an established water/solid ratio of geopolymer to the fact that it possesses a high number of acid sites as a result of
due to water evaporation (as hydration is an exothermic reaction), dealuminated framework. Mordenite, in turn, can be partially dis-
and its entrapment in zeolite pores and channels. Zeolite addition solved in high-alkaline systems as it has wider pores and, in
(22e27 wt%) caused a decrease in compressive strength in com- consequence, fewer reactive sites. Having much higher SiO2/Al2O3
parison with reference geopolymers (without zeolite) e e.g. from ratio than mordenite, zeolite ZSM-5 is more stable at high pH.
27 to 17 MPa (samples activated with potassium hydroxide and Alkali-activation of waste fluid cracking catalysts was performed by
silicate solution) or from 7.5 to 3 MPa (samples activated with Silva et al. (2016) as well. Xu et al. (2016) prepared metakaolin-
NaOH). Such mechanical strengths are sufficient for the application based geopolymers with the addition of strontium-loaded zeolite
as self-supporting monoliths. Mineralogical characterization A. In comparison to Portland cement-based products, the geo-
revealed that zeolite Na13X was present in geopolymer-zeolite polymer solidification materials exhibited higher compressive
composites. Moreover, in the case of samples activated with strength, thermal stability as well as Sr leaching resistance.
NaOH two zeolitic phases were detected e zeolite Na13X used as a
filler and zeolite A as a result of a partial transformation of geo- 2.2. Direct zeolites formation in the matrix of geopolymer
polymeric gel. This sample had the highest total pore volume and
specific surface area, and thus was most promising for CO2 The utilization of zeolites as raw materials for the production of
adsorption which had a value of 3.1%. The comprehensive study on geopolymer-zeolite composites is, however, cost generating since
carbon dioxide adsorption of these geopolymers (22e36 wt% of zeolites can form in geopolymer matrix by themselves. Their for-
zeolite Na13X) were also conducted (Minelli et al., 2018). The CO2 mation is often an accompanying reaction to geopolymerization,
capacity of Na-activated geopolymer-zeolite composite was close to and the type and amount of zeolite depend on the chemical
the values obtained for pure zeolite Na13X. In the study of Huang composition of raw materials, the type of alkaline activator as well
et al. (2012) zeolite 5 A was used as a filler in fly ash-based as curing conditions (Table 2).

Table 2
Zeolites formed in geopolymers depending on synthesis and curing conditions.

Synthesis conditions Curing conditions Obtained zeolitic structure types Ref.

Precursor Activator Si/Al Na/Al LTA FAU CHA SOD other

MK SH 1.0 1.0 60  C, 6h þ Wan et al. (2017b)


MK SH 1.0 e 50  C, 24 h þ Zibouche et al. (2009)
MK SH 1.0 0.88 80  C, 24 h þ Krol et al. (2016)
MK SH 1.0 1.0 95  C, 24 h þ Krol et al. (2016)
MK SH 1.0 1.0 95  C, 24 h þ þ Krol et al. (2017a)
MK SH 1.0 1.0 80  C, 24 h þ Krol et al. (2017a)
MK SH 1.0 1.10 40  C, 54 h þ Zhang et al. (2012)
MK SH 1.0 1.47 30  C, 72 h þ Zhang et al. (2012)
MK SH 1.03 0.85 60  C, 24 h þ Li et al. (2017)
MK SH, WG 1.13 1.0 95  C, 24 h þ Krol et al. (2017a)
MK SH 1.12 1.492 60  C, 24 h þ þ Ozer and Soyer-Uzun (2015)
MK SH 1.12 1.492 60  C, 7d þ Ozer and Soyer-Uzun (2015)
MK SH, WG 1.25 1.0 90  C, 48 h þ Rasouli et al. (2015)
MK SH, WG 1.11 1.0 95  C, 24 h þ þ Krol et al. (2016)
MK SH 1.12 1.0 90  C, 48 h þ Rasouli et al. (2015)
MK SH, WG 1.13 1.0 80  C, 24 h þ Krol et al. (2017a)
MK SH, WG 1.24 1.0 95  C, 24 h þ Krol et al. (2016)
MK SH, WG 1.25 1.0 95  C, 24 h þ Krol et al. (2017a)
MK SH 0.25 0.66 40  C, 1 h; 90  C, night þ Fletcher et al. (2005)
MK SH 0.5 0.66 þ Fletcher et al. (2005)
MK SH, WG 2.50 1.67 40  C, 53 d þ De Silva and Sagoe-Crenstil (2008)
MK SH, WG 1.9 0.99 110  C, 1 m ANA Lancellotti et al. (2013)
RHA SA 1.5 1.27 40  C, 21 d þ þ Hajimohammadi and van Deventer (2017)
GSb SA 1.5 1.3 40  C, 21 d þ þ ANA Hajimohammadi et al. (2011a)
FA SH 2.78 1.02 80  C, 6 h CAN Oh et al. (2010)
FA SH 2.3 1.31 20  C, 0.5 h CAN Sun and Vollpracht (2018)
FA SH, SA 1.0 1.0 70  C, 24 h þ þ Temuujin et al. (2010)
MSa SA 1.0 0.98 80  C, 3 d þ þ þ Sturm et al. (2016)
MSa SA 3.0 0.98 80  C, 3 d þ þ Sturm et al. (2016)
a
Microsilica.
b
geothermal silica.
562 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

The role of Si/Al ratio. One of the most important factor affecting and alkali aluminate solutions. At constant activator concentrations
the possibility of zeolites appearing during geopolymerization is hydroxysodalite was eager to form at low Si/Al ratio of 1, faujasite at
the molar ratio of silicon to aluminum (Si/Al) in a reaction system e 2 and 3, while higher ratios resulted in completely amorphous
the elements which also form zeolitic frameworks. The composi- products. Hajimohammadi and van Deventer (2017), in turn,
tion of zeolites is generally close to the composition of the gel from identified zeolite X and zeolite A in geopolymers prepared from rice
which they crystallize (Panagiotopoulou et al., 2015). Metakaolin is hull ash and sodium aluminate at Si/Al of 1.5 and H2O/Na2O of 14,
often treated as a model geopolymeric system due to its relatively while at higher Si/Al (2.5) and lower H2O/Na2O (12) no zeolites
simple chemistry (van Deventer et al., 2007). Moreover, zeolite were observed. Rice husk (or hull) ash (RHA) is an industrial waste
formation is commonly observed in metakaolin-based geopolymer generated as a by-product of burning rice husks, an abundantly
matrix. Wan et al. (2017a) identified zeolite A only at Si/Al ratio of 1 available by-product of rice milling. The worldwide production of
in such geopolymers. At extended Si/Al ratios, with silica fume as rise husks is about 70 million tons per year (Hwang and Huynh,
additional silica source, zeolites did not appeared, suggesting that 2015). The zeolitic phases were observed after 21 days of curing
higher ratios (from 1.5 to 4) prevent zeolites from crystallization. (no peaks were registered after 2, 6 and 14 days) (Hajimohammadi
Similar observations at Si/Al close to 1 was presented in the studies and van Deventer, 2017), which is related to the fact that the growth
(Zibouche et al., 2009; Kro l et al., 2016). Zeolites were more likely to of zeolite crystals is relatively slow (Komljenovi c et al., 2010) and
arise when the content of aluminum is higher, so at the lower Si/Al depends for example on silica availability. The influence of silica
ratio, confirming the optimal value of 1.0 for the formation of availability on geopolymerization was examined with the use of
zeolite A. Zeolite X appeared at Si/Al ratio higher than 1.1 and H2O/ geothermal silica (Hajimohammadi et al., 2011a). The sample with
Na2O ratio seemed not to have any significant effect on the type and the lowest Si availability needed the longest time to develop the
quantity of zeolitic products (Kro l et al., 2016). The suggestion that zeolitic phases: after 4 days of curing zeolite A was present in all
Si/Al ratio a little higher than one favors the formation of faujasite samples, while faujasite in samples with higher silica availability. In
structure is consistent with works (Buchwald et al., 2011; Kro l et al., the sample with the lowest Si availability faujasite was observed
2017a). In contrast, Rasouli et al. (2015) observed zeolite A at after 21 days of curing, so similar chemical changes were taking
slightly higher Si/Al of 1.25, while zeolite X at 1.12. In fact the place regardless silica availability, but with different rates.
reacting gel could have a lower or higher Si/Al ratio at least locally Sturm et al. (2016) obtained geopolymer-zeolite composites
where the particular zeolites were forming. When the content of using different sources of silica (microsilica and residue from the
aluminum is too high, the formation of zeolite A is hampered, chlorosilane production) and sodium aluminate. For microsilica-
though. In high-alumina metakaolin-based geopolymers with SiO2/ based samples zeolite A and hydroxysodalite were dominant pha-
Al2O3 ratios of 0.5 and 1.0 only hydroxysodalite was identified ses. However, the higher the silica content, the higher the relative
(Fletcher et al., 2005). This phase was also observed at high Si/Al amount of zeolite A and lower of hydroxysodalite. Moreover, in the
ratio of 2.46 (Kro l et al., 2016). low-silica systems also chabazite was present. For chlorosilane
In the case of fly ash-based geopolymers zeolite chabazite has residue-based specimens zeolite A was a major phase, but in low-
been observed in several studies (Temuujin et al., 2010; Azevedo silica samples hydroxysodalite, faujasite and zeolite EMT (hexago-
and Strecker, 2017; Ferna ndez-Jime nez et al., 2006; Ye and nal zeolite Y) were detected as well. It was assumed that the direct
Radlin ska, 2016; Provis et al., 2009; Bhagath Singh and formation of hydroxysodalite and early transformation of zeolite A
Subramaniam, 2017; Criado et al., 2007; Oh et al., 2014). It was into hydroxysodalite structures were caused by a higher pH.
stated that higher contents of silica favors the early formation of The role of an activator. The role of an alkali-activator in the
chabazite (Criado et al., 2007). Oh et al. (2014) obtained Al-rich Na- process of zeolites formation is rather complex. Sodium hydroxide
chabazite and Na-chabazite (herschelite) at Si/Al ¼ 2.6 and Ca/ is a commonly used alkali-activator, while potassium hydroxide is
Si ¼ 0.4 in geopolymer based on fly ash activated with NaOH at used quite rarely. They induce the appearance of different zeolitic
60  C. Zeolite A formed at Si/Al of 3.0, which is relatively high phases in a geopolymeric matrix. The concentration is an important
compared to metakaolin-based geopolymers, and Ca/Si of 0.1. factor as well. When sodium silicate solution (water glass) is used at
Chabazite zeolites possess double 6-membered rings and belong to the same time as hydroxides, the effect of introduced silicon should
ABC-6 family and the geopolymer consisting of these zeolites be also considered since Si/Al ratio is, as mentioned above, a crucial
achieved higher compressive strength since geopolymeric gel is factor. For instance, Li et al. (2017) observed the formation of zeolite
structurally related to this particular type of zeolite framework A in metakaolin-based geopolymer activated with NaOH solution,
(ABC-6). but no zeolites with water glass. It can be then supposed that Si/Al
Rees et al. (2007) investigated the effect of different silicate raised to too high level, hampering the possibility of zeolite A
concentrations on fly ash-based geopolymers (Si/Na ¼ 0.75) cured crystallization. Similar observations were noticed in the case of fly
at 30  C for 100 days. Hydroxysodalite appeared at low silicate ash-based geopolymers: sodium chabazite formed in fly ash acti-
concentrations (0e0.6 M SiO2), faujasite at moderate concentra- vated with NaOH (10 M), while no zeolites when sodium silicate
tions (0.8e1.2 M) and no zeolites at high concentrations (>3.5 M). was used as an activator (Zhang et al., 2008). The formation of
Irfan Khan et al. (2015) also found that the formation of hydrox- cancrinite and hydroxysodalite in fly ash/waste glass geopolymers
ysodalite is favored by low Si/Al ratio (1.78) and Na/Al ratio higher activated with sodium hydroxide was showed in the study
than 0.8. Generally, hydroxysodalite is an easily-forming phase in (Bobirica et al., 2015). In the case of activation with sodium silicate
the case of fly ash-based geopolymers (Azevedo and Strecker, 2017; no zeolites crystallized. The presence of sodium silicate can affect
Fernandez-Jime nez et al., 2006; Provis et al., 2009; Criado et al., not only the possibility, but also the type of zeolites forming in
2007; Bo €ke et al., 2015; Temuujin et al., 2014). Komljenovi c et al. geopolymeric matrix. Moon et al. (2014) activated basaltic-type
(2010) activated different fly ashes with various activators and natural pozzolanic material in order to obtain geopolymers:
showed that the obtained products were mainly amorphous and zeolite Y and hydroxysodalite were formed in samples activated
the formation of zeolite faujasite was rather the exception than the with sodium hydroxide, while phillipsite and small amount of
rule. The increase in silicon ions content in the solution resulted in zeolite Y in the case of sodium hydroxide/water glass activation.
the decrease in the probability of zeolite formation. Similar obser- Panagiotopoulou et al. (2015) stated that the increase in silicon
vations were presented by Buchwald et al. (2011) in the case of content in the activator solution leads to the formation of more
aluminosilicate geopolymeric gel prepared from pure alkali silicate disordered phases. At the lowest Si/Al ratio (1.0) hydroxysodalite
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 563

was identified, at Si/Al ¼ 2.4 faujasite, while at the highest ratio no L, confirming that the activation with KOH is less related to zeolite
zeolites were detected. The use of activation solutions with low formation (Hu et al., 2017). However, it positively affects the for-
silicon and high alkali content favors the formation of zeolitic mation of an amorphous geopolymeric network (Panagiotopoulou
phases. Ma et al. (2016) observed the transition of geopolymeric gel et al., 2015). The presence of zeolite KeI (K2Al2Si2O8$3.8H2O) in
into zeolite (analcime) in fly ash-based geopolymers activated with blast furnace slag activated with KOH (4 and 8 M) as well as potas-
sodium hydroxide already at 25  C, while with sodium silicate only sium gismondine (in the case of 4 M KOH) was observed in the study
at 80  C, which was related to the mechanism that soluble silicates (de Moraes Pinheiro et al., 2018). The effect of olive-stone biomass
hamper the crystallization of zeolites in geopolymers. Torres- ash was also investigated: it hampered the formation of zeolitic
Carrasco and Puertas (2015) examined glass waste as a potential phases since zeolite KeI was not detected. In geopolymers based on
activator (accompanied by NaOH solution) in fly ash geopolymer metakaolin faujasite and zeolite V (Na3K3Al3Si3O24$12H2O) were
systems. Hydroxysodalite and Na-chabazite as well as N-A-S-H gel found (Duxson et al., 2007). Analcime, zeolite NaeP1 and two zeo-
were detected in the obtained materials. The same was observed in lites with natrolite type structure: gonnardite ((Na,Ca)2(Si,A-
conventional systems, activated with NaOH and NaOH/water glass. l)5O10$2H2O) and mesolite (Na2Ca2Al5Si9O30$8H2O) were detected
In the case of the concentration of alkalis, it was observed that in fly ash-based geopolymers activated with sodium silicate
the formation of zeolite A in metakaolin-based geopolymers is (Rodríguez et al., 2013b). The activation with nanosilica resulted in
preferred at lower alkali content (10 M vs. 12 M NaOH) and at the formation of faujasite with no other zeolitic phases. The acti-
higher temperature, since zeolites prefer to form at slightly lower vation with potassium silicate, in turn, caused the crystallization of
alkalinity and higher reaction temperature (high alkali concentra- K-chabazite. Elert et al. (2015) treated clay with 5 M solution of NaOH
tions favor hydroxysodalite formation rather than zeolite A) (Zhang and KOH and investigated phase composition after 1 week,1, 2.5, 4, 6
et al., 2012). The similar observation was presented in the study months and one year of treatment. In the case of NaOH treatment,
(Boonjaeng et al., 2014), where zeolite A formed at extremely low faujasite was observed after 6 months and hydroxysodalite after a
concentrations of activator (0.01 and 0.1 M NaOH) in the matrices of year. The KOH environment induced the formation of zeolite KeI and
alkali-activated metakaolin with Ca(OH)2 (metakaolin/calcium hy- chabazite after 4 months and 1 year, respectively. The formation of F-
droxide mass ratio of 0.4). The dependence of zeolite type on sys- type zeolite (KAlSiO4$1.5 H2O) in metakaolin foams activated with
tem alkalinity is related to the fact that some building units of potassium silicate was observed in the works (Prud’Homme et al.,
zeolites remain stable under different alkalinities. At higher con- 2011; Prud’Homme et al., 2012; Delair et al., 2012).
centrations (3 and 5 M), chabazite and zeolite ZK-14 were identi- The formation of natrolite was proposed to improve the strength
fied. The fact that increasing system alkalinity (NaOH of BFS-FA geopolymeric binders (Huang et al., 2017). It was also
concentration) favors the formation of zeolite ZK-14 was also re- detected in the same system also with the addition of metakaolin
ported in activated calcined kaolinitic-illitic clay (El Hafid et al., (Huang et al., 2018), while in geopolymers based on heated

2017). Alvarez-Ayuso et al. (2008) prepared fly ash-based geo- kaolinitic-illitic clay also chabazite was present (El Hafid and
polymers and observed that chabazite and hydroxysodalite were Hajjaji, 2015). In fly ash-based geopolymers natrolite was detec-
present in most cases, zeolite X and zeolite 4 A in geopolymers ted together with zeolite NaeP1 (Bhagath Singh and Subramaniam,
obtained at low alkaline solution (5 M NaOH), while zeolite NaeP1 2016). At higher SiO2/Na2O ratio (2.16) natrolite and zeolite NaeP1
at higher alkalinity (8 and 12 M). In the case of aluminosilicate were identified, but the latter disappeared after 7 days. At lower
geopolymeric gel prepared from pure alkali silicate and alkali ratio (1.51) the same zeolites were detected, however, the amount
aluminate solutions, at Si/Al ratio of 1.0, the pH was a crucial factor of zeolite NaeP1 was increasing with time. Natrolite has been also
in terms of zeolites crystallization (Buchwald et al., 2011). The found in circulating fluidized bed (CFB) combustion fly ash geo-
lower the concentration of NaOH, the higher the formation ability polymers (Eiamwijit et al., 2015), also when silica fume was added
of zeolite A and zeolite X. A high NaOH concentration favored (Chindaprasirt et al., 2014), which was attributed to a high con-
hydroxysodalite crystallization. A similar conclusion was drawn in centration of activator (15 M NaOH). Also zeolite Y formed in such
the studies (Rozek_ et al., 2018; Krol et al., 2017b), where the in- geopolymers with increasing peaks intensity with decreasing SiO2/
tensity of diffraction peaks corresponding to hydroxysodalite was Na2O molar ratio (increasing alkalinity). Tyni et al. (2014), in turn,
growing with the increasing content of NaOH in the reaction me- identified faujasite and Na-chabazite in geopolymers based on CFB
dium of fly ash-based geopolymers. Moreover, the addition of so- bottom ash. In the case of samples prepared with 10 M NaOH
dium aluminate promoted the formation of hydroxysodalite (Kro l activator solution, both zeolitic phases were observed, while in
et al., 2017b). geopolymers prepared with 5 M NaOH only chabazite was detected.
The type of zeolite that forms parallel to geopolymerization de- The role of curing. It should be noted that curing of geopolymers
pends on the dominant cation in an activator solution. Shearer et al. is directly related, e.g. with alkalinity and silica to alumina ratio. It
(2016) prepared geopolymers based on fly ash from co-combustion was observed that different temperatures of activation result in the
of coal and biomass and detected in their matrices Na-chabazite, formation of various zeolitic phases. Cwirzen et al. (2014) activated
faujasite and analcime (activation with 7% Na2O). An increase in metakaolin and limestone blends with NaOH (3 M), which resulted
activator content reduced the formation of zeolites, especially fau- in the formation of zeolite A, both at low (20  C) and high (80  C)
jasite. Also, an increase in soluble silica slowed their crystallization. temperature of curing. However, the activation with 5 M NaOH
For ashes activated with potassium no zeolites formed, suggesting showed a significant temperature effect, i.e. no zeolites at a low-
that the formation of potassium zeolites is not so favored as the temperature activation, and faujasite formation at a higher tem-
formation of sodium ones. The kind of alkali ion affects not only the perature. In the case of Si/Al different zeolitic phases were obtained
dissolution of the aluminosilicate material, but also the poly- in the matrix of metakaolin-based geopolymers depending on
condensation of a geopolymeric gel (Kro l et al., 2018). Bakharev curing temperatures as well: at Si/Al of 1.12 zeolite A at 95  C and
(2005b) observed zeolite NaeP1 and hydroxysodalite in fly ash- zeolite X at 80  C, while at 1.25 zeolite X at 95  C and no zeolites at
based geopolymers. These phases were observed to be more or- 80  C (Kro
l et al., 2017a). Granizo et al. (2014) evaluated time and
dered in the case of the sample activated with NaOH than in the temperature of metakaolin-based geopolymers curing (Fig. 3). Ze-
sample activated with NaOH together with KOH. The activation of olites started to form after 10 h of curing. Zeolite A and zeolite X
meta-illite-smectite clay with NaOH at 75  C resulted in the for- formed at 25  C after 9 days and hydroxysodalite was detected in
mation of phillipsite, while the use of KOH gave only traces of zeolite 36-day sample. At 85  C reactions were much faster, namely after
564 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

Fig. 3. Zeolites present in the solid phase at (a) 25 and (b) 85  C (N5 e activation with 5 M NaOH, N8 e 8 M NaOH). Reproduced with permission from Ref. (Granizo et al., 2014).
Copyright 2014 Elsevier.

8 h zeolite A began to transform into hydroxysodalite and to after 7-day curing. The phase evolution in geopolymers based on
disappear totally after 14 days. Zeolites are metastable and may geothermal silica and sodium aluminate, seeded with 0.5 wt% of
successively transform into one or several more stable phases oxide (Al2O3, ZnO and ZrO2) nanoparticles, was observed in the
(Criado et al., 2007). Also, an increase in the crystallinity of zeolite work (Hajimohammadi et al., 2011b). Zeolite A was present in the
A, faujasite and hydroxysodalite in geopolymers based on meta- samples after 4 days at 40  C, while after 3 weeks zeolite A as well
kaolin and rise husk ash was observed with increasing curing as faujasite and analcime. After 8 months faujasite peaks became
temperature (Sore et al., 2016). In geopolymers based on water stronger in XRD patterns of seeded samples, while disappeared in
treatment sludge and fly ash chabazite formed at 65, 75 and 85  C, the case of unseeded sample since seeding reduces the nucleation
while gismondine only at higher temperatures (75 and 85  C) time required for zeolite synthesis. Criado et al. (2007) concluded
(Suksiripattanapong et al., 2015). In the case of geopolymers pre- that the formation of well-crystallized zeolites requires a long
pared with using of sewage sludge ash as a partial replace of period of time, while amorphous zeolite precursor gel is expected
metakaolin, zeolite (faujasite) crystallized in the samples cured at to form at ambient conditions in a short reaction time. The trans-
65  C in a thermal bath, while no peaks were observed in case of formation of amorphous phase into chabazite or zeolite X into ze-
25  C (Istuque et al., 2016). The effect of curing regime on the type olites A and P in metakaolin-based geopolymers was observed
of zeolites formed in fly-ash based geopolymers activated with between 53 and 210 days of curing at 40  C and corresponding
NaOH was shown in the study (Bakharev, 2005a). Zeolite A and strength reduction (De Silva and Sagoe-Crenstil, 2008). Such phe-
chabazite were present in the samples precured in room temper- nomenon was also observed in the case of commercial X-type
ature for 2 h and then cured at elevated temperature (75  C) for one zeolite that transformed into the more stable P-type zeolite and
month, while chabazite, zeolite NaeP1 and hydroxysodalite in the hydroxysodalite (Takeda et al., 2013). Analcime was observed in
samples with precuring time of 24 h and 6 or 24 h of heat curing. metakaolin-based geopolymer cured at a relatively high tempera-
Duan et al. (2015) used metakaolin as a precursor e zeolite A ture of 110  C for 1 month (Lancellotti et al., 2013). It is the most
occurred in the matrix of geopolymers cured for 16 h at 70  C and stable and densest common zeolitic phase (Criado et al., 2007). The
higher temperatures, however, at 110 and 130  C peaks corre- extended curing time enhanced the formation of chabazite and
sponding to zeolite A were significantly weaker. The highest gismondine in silty clay-fly ash geopolymers (Sukmak et al., 2013).
amount of zeolite A was observed in geopolymer synthesized with In the study (Criado et al., 2010), the quantity of zeolites (chabazite,
Na/Al ratio of 1.0, water/binder ratio of 0.7, cured at 70  C for 16 h. hydroxysodalite) was increasing with time as follows: 4.9% (8 h),
Bernal et al. (2011) observed the formation of gismondine in 24.8 (7 d), 28.1 (28 d), 36.5 (180 d).
alkali-activated granulated blast furnace slag and its blends with The role of raw materials, additives and other factors. Slaty et al.
metakaolin. An increase in curing times contributed to the forma- (2013) observed the formation of Na-zeolites (phillipsite and
tion of zeolitic phases. The presence of gismondine was not stated natrolite) and hydroxysodalite in kaolinitic clay activated with
at early ages of curing, suggesting that longer time of curing is NaOH at 80  C for 24 h. The use of raw kaolinitic clay, however,
required for the development of full crystallinity in this phase as a results in the presence of unreacted phases such as kaolinite,
consequence of its complex structure. Ozer and Soyer-Uzun (2015) muscovite and illite in the matrix of geopolymer. Heah et al. (2012)
observed that the time of thermal curing (at 60  C) influenced the activated kaolin with NaOH and WG and observed the formation of
formation of zeolite phases in metakaolin-based geopolymers zeolite X after curing at 80  C for 24 h. In the case of metakaolin-
(with Si/Al ratio of 1.12): mainly zeolite A and small amount of based geopolymers, the temperature of kaolinite calcination may
hydroxysodalite were present after 1-day, while mainly the latter be an additional factor affecting the formation of zeolites. Glid et al.
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 565

(2017) showed that zeolite A formed in shorter times in the case of


kaolinite heated in 550 than in 800  C, namely 20 min versus
120 min. It was observed that the quantity of zeolite increased with
increasing specific surface area of metakaolin (Fig. 4) e the higher
the specific surface area, the higher the solubility of Al and Si ions,
resulting in an increase in zeolite formation (Takeda et al., 2013).
The activation (9 M NaOH at 70  C) of metakaolin obtained by
kaolinite calcination at 700  C resulted in the formation of zeolite A,
while from metakaolin calcined at 800  C only amorphous matrix
was obtained (Kro l and Rozek,
_ 2019). Belviso et al. (2015) prepared
geopolymers on the basis on fly ash/kaolinite mixtures. The tem-
perature of activation was relatively low, 45  C, and time differed
from 1 to 216 h. Zeolites X and ZK-5 were detected in the geo-
polymer based only on fly ash. The increasing part of kaolinite
resulted in a decrease in zeolite X amount, an increase in zeolite A
and a total elimination of zeolite ZK-5. FAU-type zeolites have been
Fig. 4. Relationship between the specific surface area of metakaolin and the amount of
also detected in geopolymers based on fly ash (Ye and Radlin  ska,
zeolite in final materials (Takeda et al., 2013).
2016) and rice husk ash (Hwang and Huynh, 2015), in low-silica
fly ash-based geopolymers (Provis et al., 2009), and metakaolin-
based geopolymers (Takeda et al., 2013; Liew et al., 2012). kaolin blends activated with sodium silicate and sodium or potas-
Peng et al. (2017a) observed zeolite NaeP1 and hydroxysodalite sium hydroxides. Hydroxysodalite was not detected when meta-
in a one-part geopolymer obtained by alkali-fusion of bentonite kaolin was used instead of kaolin, as kaolin provides OH ions
with 20% NaOH at 700  C. Faujasite and hydroxysodalite, in turn, which accelerates the reaction of kaolin with NaOH, while meta-
appeared when Na2CO3 was used as an activator at 850  C and also kaolin does not contribute to the formation of hydroxysodalite
zeolite NaeP1 at 1000  C. This zeolite was also found in a one-part because it is the product of dehydroxylation. Zeolite ZK and
geopolymer obtained by hydration of bentonite calcined together hydroxysodalite were observed in fly ash-based geopolymers
with dolomite and Na2CO3 (Peng et al., 2017b). Naghsh and Shams activated with 8 M solution of NaOH (zeolite ZK is similar to so-
(2017) applied a two-step method of geopolymer preparation, dalite) (Nath et al., 2016). Hydroxysodalite has been also identified
including a fusion of kaolin with sodium hydroxide (400e800  C) in geopolymers based on, e.g. FA with BFS (Bobirica  et al., 2015;
and hydration/dealkalization (washing with distilled water), so Gordon et al., 2014; Muhammad et al., 2018), coal bottom ash
kaolin may be used directly, without the necessity of its trans- (Geetha and Ramamurthy, 2013), bagasse bottom ash (waste of
formation into metakaolin. Zeolite ZSM-5 was detected in geo- sugar industry) and natural clay (together with chabazite) (Noor-
polymer obtained from fusion at 800  C for 14 h with NaOH/kaolin Ul-Amin et al., 2016) as well as in alkali-activated aluminosilicate
ratio of 2.2, while at 600  C geopolymer was completely amor- glass with various CaO content (Garcia-Lodeiro et al., 2016). In
phous. The presence of zeolite Y in one-part-mixing geopolymers geopolymers based on heated clay with milled waste glass also
based on metakaolin geopolymer powder was confirmed as well chabazite was found (El Hafid and Hajjaji, 2018). Also, zeolites of
(Yun-Ming et al., 2017). gismondine-type structure (GIS) have been mentioned in several
Ye et al. (2014) prepared geopolymers based on Bayer red mud works concerning geopolymers: garronite and zeolite NaeP1 were
(waste residue from caustic soda digestion of alumina from bauxite detected in geopolymers based on slag/fly ash (25 wt%/75 wt%) and
ore in the Bayer process). Cancrinite zeolite crystallized in the fly ash (Ismail et al., 2014), thomsonite and garronite in MK/BFS-
samples containing only red mud as well as containing red mud based geopolymers (Zhang et al., 2010), gismondine in Ca-rich fly
and granulated blast furnace slag (RM/GBFS 1:1 wt/wt). Its forma- ash-based geopolymer activated with mixed sodium hydroxide and
tion was related to adsorption of carbon dioxide from air in a highly sodium silicate (Guo et al., 2010a) and in complex geopolymers
basic environment according to the equation: based on coal fly ash and other Ca-containing materials (Portland

2CO2 þ 6Naþ þ 2Ca2þ þ 6H3 SiO


4 þ 6AlðOHÞ4 /Na6 Ca2 Al6 Si6 O24 ðCO3 Þ2 , 2H2 O þ 18H2 O þ 2OH


Cancrinite was also observed in one-part geopolymer binders cement, flue gas desulfurization gypsum, and water treatment re-
prepared from red mud with the addition of silica fume (Ye et al., sidual) (Guo et al., 2010b), zeolite NaeP1 in the samples of meta-
2016), geopolymers based on pumice-type natural pozzolanic kaolin activated with sodium silicate solution and cured at 80  C in
material (together with chabazite) (Najafi Kani et al., 2012), NaOH- water bath (Zhu et al., 2013) as well as in halloysite-based geo-
activated fly ash (Sun and Vollpracht, 2018; Oh et al., 2010), BFS/FA- polymer (Takeda et al., 2013).
based geopolymers (Xu et al., 2014), and geopolymers based on The presence of limestone in the system of alkali-activated
waste sandstone sludge (zeolite A and sodalite formed instead metakaolin enhanced the formation of zeolites (in comparison to
when sludge was replaced by 50 wt% of metakaolin) (Clausi et al., samples based only on metakaolin) (Cwirzen et al., 2014). More-
2018). Zeolite vishnevite (Na8Al6Si6O24(SO4)$2H2O, CAN structure) over, the higher content of limestone favored the formation of
was observed in geopolymers based on bottom and fly ash from hydroxysodalite, while the lower content favors zeolite A. Jun and
pulverized coal combustion (Boonserm et al., 2012). Oh (2014) observed Al-rich chabazite, zeolite A and zeolite X in
Komnitsas et al. (2009) observed the formation of hydrox- fly ash-based geopolymers e chabazite was found only in geo-
ysodalite in geopolymers based on electric arc ferronickel slag and polymer based on fly ash containing 10 wt% CaO, and was absent
566 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

when CaO content was lower than 5 wt%. The samples with cha- formation of hydroxysodalite and traces of zeolite A (Ferna ndez-
bazite showed higher compressive strength than the samples with Jimenez et al., 2008). Then the samples were treated at 150  C
zeolite A, although it can be related to the formation of CeSeH and faujasite and greater amount of zeolite A were detected. That
phase due to the higher content of CaO. Tzanakos et al. (2014) thermal treatment caused the increase in the compressive strength
prepared geopolymer based on metakaolin and medical waste of alkali-activated matrices. The similar situation was observed for
incineration ash (fly ash and bottom ash) and obtained faujasite in metakaolin/fly ash geopolymer. After the exposure of fly ash/slag
its matrix. The addition of calcium carbonate caused the formation geopolymer to 800  C dehydrated Ca-substituted zeolite A
of phillipsite. It was however stated that sodium aluminosilicate appeared (Zhang et al., 2015). The formation of zeolite NaeP1 was
hydrate containing calcium (Ne(C)eAeSeH) is more stable, and observed after thermal shock of fly ash-based geopolymers (Nazari
with the increase in the content of calcium and other elements, the et al., 2017).
transition from amorphous gel to crystalline zeolite becomes more The factor which can induce the formation of zeolites in geo-
and more difficult (Ma et al., 2016). polymers is pH. Nguyen and Skv   ara (2016) provided evidence for
Fernandez-Jime nez et al. (2017) alkali-activated fly ash and glass the appearance of zeolitic phase (minor amounts of zeolite NaeP1)
powder with cleaning solution from aluminum casting industry. In in the sample of fly ash-based geopolymer immersed in the solu-
both cases herschelite, zeolite NaeP1 and hydroxysodalite tion of NaOH (pH of 14) for 360 days. Shi et al. (2018) observed the
appeared in a geopolymeric matrix. These zeolites did not form formation of zeolites (zeolite A, zeolite X and zeolite NaeP1) in the
when NaOH (8 M) was used as an activator, which suggests that the case of pastes and mortars of alkali-activated slag with metakaolin
character of cleaning solution induced zeolites crystallization, exposed to 1 M NaOH. Zeolite NaeP1 also formed in pastes with fly
namely its lower alkalinity and soluble Al presence. When high ash after the same exposure. Zeolite NaeP1 and hydroxysodalite
surface area Al2O3 nanoparticles were introduced to a fly ash geo- were observed in fly ash-based geopolymers, and after the expo-
polymeric mix, zeolite NaeF (Na2Al2Si2O8$xH2O) formed e a syn- sure to sulfate solutions, chabazite was also identified (Bakharev,
thetic zeolite with the edingtonite structure type (EDI), observed in 2005b). Zeolite NaeP1 and chabazite appeared in the fly ash-
geopolymers for the first time (Rees et al., 2008). In the seeded based geopolymer activated with sodium silicate after its expo-
geopolymer system the formation of zeolite F is favored in the re- sure to 5% sulfuric acid solution (Bakharev, 2005c). The presence of
gions of Al-rich gel located close to the original nuclei, which NaSO4, in turn, in an activating solution retarded the formation of
provides the proper stoichiometry. The presence of this Al-rich gel, N-A-S-H gel in alkali-activated fly ash, but accelerated its trans-
which did not occur in the unseeded system, explains the formation formation into zeolites, such as Na-chabazite, zeolite NaeP1 and
of zeolite F at a lower Si/Al ratio than zeolite X that formed in the hydroxysodalite (Criado et al., 2010).
absence of seeding. In the case of alkali-activated Ca-rich clay with The increasing content of lead immobilized in fly ash-based
the addition of AlF3 production waste (a by-product silica with a geopolymers hampered the formation of hydroxysodalite (Lee
high AlF3 content, i.e. 60.15% and 35.52%, respectively), only et al., 2016b). In fly ash-based geopolymers immobilizing arsenic
hydroxysodalite was present in the samples with 0, 10 and 25% (1 wt%), the formation of hydroxysodalite and herschelite was not
addition of the AlF3 production waste (Vai ciukyniene_ et al., 2018). disturbed (Fernandez Jiminez et al., 2004). The formation of zeolite
Zeolite A formed as a product of hydroxysodalite recrystallization in A was shown in water treatment residue-based geopolymers, and
the geopolymer with 50% addition, which caused a slight decrease NaeZn-exchanged zeolite A in the case of zinc immobilization in
in compressive strength. Liu et al. (2017a) prepared foamed geopolymeric matrix (Waijarean et al., 2017). Bell et al. (2008)
geopolymer-zeolite blocks from fly ash with oleic acid and Al mixed metakaolin with cesium silicate solution and cured the
powder, cured at 80  C for 36 h. Faujasite was a dominant phase in mixture at 50  C for 24 h. After heating the obtained geopolymer to
geopolymer with the addition of 0.05 wt% of Al powder, while in 1100  C, the formation of pollucite was observed. Pollucite (CsAl-
the case of 0.10 wt% zeolite NaeP1 was also present as a trace. It is Si2O6), a Cs-bearing zeolitic phase, has a significant industrial in-
therefore possible to design the type of zeolite by controlling terest as an encapsulant for radioactive 137Cs as well as a refractory
aluminum content in the mixture. material, because of a high melting point and very low coefficient of
Liu et al. (2017b) prepared Al2O3eSiO2 powder by the sol-gel thermal expansion. A unit cell of pollucite is built of a network of 48
method, activated it with sodium silicate, and cured at 60  C for corner-shared SieO or AleO tetrahedra with 16 Csþ ions located in
6 h. It was shown that only the samples with H2O/Na2O molar ratio cavities, 12-coordinated to oxygen. The cavities are formed since
higher than 8 can transform a geopolymeric gel into zeolite A. The the tetrahedra are linked to form 4-, 6-, and 8-membered rings. The
limitations on zeolites formation in systems with low amount of 8-membered rings create elongated cages that permit large cations
free water are related to the fact that nutrients needed for the (Csþ and Rbþ) to pass through 6-membered rings. It is assumed that
crystallization cannot be provided due to transportation difficulties. only minor atomic rearrangements in geopolymer structure are
Additionally, the suspension solidification method was used to required for its transformation into crystalline pollucite due to the
prepare spheres e the geopolymeric slurry with H2O/Na2O ¼ 10 similarities in short and medium range order between these two
was dosed in droplets into silicon oil or polyethylene glycol (PEG- phases (Bell et al., 2008). Ofer-Rozovsky et al. (2016) activated
6000). The geopolymer spheres formed in silicone oil transformed metakaolin with the solution of NaOH and CsOH at 40  C and
into zeolite A, while in polyethylene glycol did not transform into observed the formation of Cs analogue of a zeolite F (CsAlSiO4$H2O)
zeolites. When water diffuses from the inside to the surface of a in the case of 50%-substitution of Na by Cs in an activation solution.
sphere, it is sealed due to the fact that is insoluble in silicon oil. At lower Cs content (7%) also zeolite A and zeolite X were detected,
Tetrahedra of [SiO4] and [AlO4] rearrange around free sodium ions while at the lowest (1%) only zeolite A and zeolite X. Zeolite F (EDI
because of the transport of water molecules and transform into framework) has a relatively open structure. The EDI structure has
zeolites around the formed nanocrystalline core. Since water is two sets of channels, which are comprised of bent 8-membered
highly soluble in polyethylene glycol, water molecules with sodium rings. The minimal distance between oxygen atoms is large
ions can diffuse outside a sphere, so there is no medium that can enough for the inclusion of hydrated Cs or Na cations. Shiota et al.
transport and rearrange silicate and aluminate groups in the geo- (2017a) prepared geopolymers based on municipal solid waste
polymer gel which can transform into zeolite. The mechanism is incineration (MSWI) fly ash and dehydrated pyrophyllite (a 2:1
also discussed in Section 2.4. layer aluminosilicate [Al2[Si4O10](OH)2] with repeating layers of an
Alkali-activation of metakaolin at 85  C resulted in the [AlO6] sheet between two [SiO4] sheets). Faujasite and sodalite
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 567

were found in this material also in the presence of radioactive ce- shown that, in these conditions, the increasing SiO2/Al2O3 molar
sium (0.1% CsCl). At high CsCl addition (40%) sodalite and albite ratio of geopolymer caused a gradual increase in the intensity of
formed, and in this case also pollucite crystallized. Faujasite and diffraction peaks corresponding to faujasite. At 100  C it was
sodalite (Na8Al6Si6O24Cl2) were also identified in the same MSWI possible to obtain FAU-type as well as NaeP1 zeolite phases. Lee
fly ash-dehydrated pyrophyllite system geopolymer aimed to et al. (2016a), in turn, investigated fly ash/blast furnace slag mes-
immobilize Pb (Shiota et al., 2017b). oporous geopolymer-zeolite materials. Geopolymeric pastes were
cured at 80  C for 24 h, and then hydrothermally treated at 90  C
also for 24 h. The presence of zeolite NaeP1 and hydroxysodalite
2.3. Hydrothermal treatment of geopolymers was observed in the samples with and without slag, while faujasite
and chabazite only in the sample without slag. The compressive
When geopolymerization is conducted at room temperature strength was substantially improved by the addition of slag (from
and under atmospheric pressure, water/solid ratio is often insuffi- 2 MPa for the sample without slag to 15 MPa for the sample with
cient for achieving the crystallization conditions, so geopolymers 40% of slag). Alkali-activated hydroceramics based on slag-
are maintained in a metastable state. However, when a hydro- metakaolin-fly ash systems for waste immobilization were pre-
thermal treatment (HT) is applied, the geopolymer transforms into pared by Wang et al. (2016). The materials were prepared by acti-
zeolite crystals more readily (Yan et al., 2012). This process pro- vation of aluminosilicate precursors blends, precured at room
motes the depolymerization of SieAl radical ions as well as en- temperature and hydrothermally treated (150e210  C for 12e60 h).
hances the formation of zeolite grains and supports the growth of Zeolite NaeP1 was a dominant phase in the samples treated below
crystals in geopolymer matrix (Liu et al., 2016a). It is usually carried 180  C but analcime was also present. From 180 to 210  C the
out in a Teflon-lined hydrothermal reactor (autoclave). The syn- diffraction peaks of analcime became stronger, while those of
thesis of geopolymer-supported zeolites follows a typical two-step zeolite NaeP1 weaker till to complete disappearance at 210  C. The
procedure: (1) synthesis of bulk geopolymers, and (2) nano- treatment time was also an important factor e the longer the time,
structural conversion of the part of geopolymeric gel into zeolites the greater the amount of analcime and smaller of zeolite NaeP1.
under hydrothermal conditions (Khalid et al., 2018). In a one-step De Rossi et al. (2018) prepared geopolymers based on biomass fly
procedure, in turn, geopolymeric slurries are immediately put un- ash (75 wt%) and metakaolin (25 wt%) and cured them at RT for
der hydrothermal conditions for direct nucleation of some alumi- 24 h. Then, the samples were cured in hydrothermal conditions in
nosilicate oligomers into zeolites, while the rest of oligomers hermetic glass container for 3, 7 and 28 days at 60  C. Such a
polymerize into amorphous geopolymeric gel. Then, some of the treatment caused the formation of zeolite NaeP1 and faujasite
geopolymeric gel may transform to zeolites at a later stage of hy- regardless the curing time, however, the higher time of the treat-
drothermal treatment as well (Khalid et al., 2018). A schematic ment (28 days) enhanced their formation. Tang et al. (2016) syn-
diagram of these procedures is presented in Fig. 5. This section thesized zeolite ZSM-20 in two-part procedure, i.e. the synthesis of
summarizes different approaches made to obtain zeolitic phases metakaolin-based geopolymer (Si/Al ¼ 2.0, activation with sodium
via hydrothermal treatment of geopolymers. silicate solution with modulus of 1.1, H2O/Na2O molar ratio of 7.5,
Qiu et al. (2014) synthesized fly ash-based geopolymers (curing curing at 40  C for 3 days), and its hydrothermal treatment (140  C,
at 90  C for 24 h) and treated them hydrothermally (80e110  C for 10 h). ZSM-20 zeolite resembles faujasite in certain structural as-
6e48 h) in order to obtain blocks with zeolite NaeP1. Zeolite pects, but it has a higher Si/Al ratio. The hexagonal ring remains
formed regardless SiO2/Al2O3 ratio (1.0e2.2) of geopolymers, but stable in low-alkalinity media, so the synthesis of ZSM-20 should
with the best crystallinity at 1.4. Zeolite NaeP1 was obtained even if be conducted under appropriate conditions. Different conditions
deionized water was used as a reaction solution during HT, but allowed to obtain other zeolites such as analcime, faujasite, zeolite
NaOH solution (1.0, 2.0, 2.5 M) gave greater amounts of a zeolitic NaeP1, and hydroxysodalite. Kolousek et al. (2007) applied hy-
phase. At the highest alkalinity of the solution (3.0, 4.0 M) also drothermal conditions (100 or 140  C for 1e7 days) for metakaolin-
hydroxysodalite was present in the samples. The optimal temper- based geopolymers. Such treatment enhanced the reactivity of
ature of hydrothermal treatment was 100  C, while time 24 h. Liu quartz, and caused the formation of chabazite and phillipsite in
et al. (2016a) obtained a faujasite block from a fly ash-based geo- short-term period (1e4 days) or analcime and hydroxysodalite in
polymer as well. The geopolymeric slurry was molded and cured in long-term period (7 days, 140  C). Geopolymers synthesized from
air for 12 h and in 80  C for another 12 h. The obtained samples Al2O3eSiO2 powders fabricated via sol-gel method, which allowed
were hydrothermally treated (60e100  C, 1e24 h) with water or to obtain geopolymers with low Si/Al ratio and to overcome the
NaOH solution. The optimal conditions were: 70  C, 24 h, 1 M NaOH disadvantages related to the use of impure raw materials, were,
solution, which gave a compressive strength of 5 MPa. It was also

Fig. 5. Schematic diagram of geopolymer-zeolites composites preparation process via hydrothermal treatment.
568 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

after curing at 60  C for 24 h, treated hydrothermally in Teflon phase, but increased compressive strength to 20 MPa (in compar-
bottles with water at 90  C for 6 h, which caused the formation of ison to 12 MPa for the sample without slag). Zeolite NaeP1 spheres
faujasite (at SiO2/Al2O3 ratio of 3) and zeolite A (at SiO2/Al2O3 ratio were prepared via hydrothermal treatment of metakaolin-based
of 2) (Yan et al., 2012; Xuemin et al., 2011). Ge et al. (2014) applied geopolymer spheres (Tang et al., 2015). Activator (sodium silicate)
the geopolymer-gels-conversion method to obtain large-sized was added to metakaolin together with sodium dodecyl sulfate
analcime. The chemosynthetic geopolymers were prepared as (K12) in order to obtain foamed slurry, which was dosed in droplets
described above, cured at 70  C for 72 h, and hydrothermally to polyethylene glycol (PEG-600) in water bath (70  C). The droplets
treated at 240  C for 6 h. It was shown that the grain size of anal- were suspended and retained a spherical shape due to the fact that
cime can exceed 1 mm when the time of hydrothermal treatment the densities of the droplets and PEG-600 were approximately
exceeded 48 h at 240  C. equivalent. The separation of the spheres from the PEG-600 me-
Zhang et al. (2014) prepared a self-supporting zeolite X mem- dium was facilitated by their rapid solidification. The spheres were
brane in a form of a thick film (10 mm thickness). The geopolymer then cured at 60  C for 3 days. The final step was hydrothermal
slurry was cured at 60  C for 24 h, then solidified geopolymers were treatment at 140  C for 10 h.
treated hydrothermally at 90  C for 9e15 h. It was shown that by Khalid et al. (2018) applied hydrothermal treatment (100 and
extending the hydrothermal treatment time, the relative crystal- 125  C for 24 and 48 h) for the preparation of fly ash/slag
linity and surface density of the membrane can be enhanced. Blast geopolymer-supported zeolites using one-step method (without
furnace slag was used by Azarshab et al. (2016) as a precursor pre-curing). The major formed crystalline phase was zeolite NaeP1,
material for geopolymer-zeolite membrane synthesis. The pre- while Na-chabazite, analcime and hydroxysodalite were present as
pared geopolymeric paste (slag activated with NaOH solution) was minor phases in the samples containing 80 wt% of FA and 20 wt% of
formed into discs (diameter of 25 mm, thickness of 3 mm), kept at BFS. The increase in the content of slag (to 60 wt%) caused a sig-
ambient temperature for 24 h, and then hydrothermally treated at nificant decrease in the intensity of zeolite NaeP1 peaks, and
90  C for another 24 h. Hydrothermal treatment resulted in the hydroxysodalite became a major phase. This was related to the
transformation of geopolymer, which was mainly amorphous with presence of Ca in slag, which has a tendency to hamper the for-
a small amount of metaheulandite zeolite (CaAl2Si7O18$3.5H2O, mation of zeolites. Ca also contributes to the creation of C-A-S-H
HEU framework), into a geopolymeric membrane coated with gel, which is why this sample achieved the highest compressive
yagawaralite zeolite (CaAl2Si6O16$4H2O, UYG framework) layers of strength (ca. 25 MPa). The specimens with 20 wt% of BFS had a
5 mm thickness on both sides. It was stated that metaheulandite slightly lower strength, i.e. between 15 and 20 MPa. Wang et al.
zeolite in hardened geopolymer gel acted as initial seeds for the (2018) presented one-step geopolymer-gels-conversion of pre-
formation and growth of yugawaralite zeolite in the membrane. shaped (by curing in oven for 30 min) metakaolin-based geo-
Self-supported ECR-1 zeolite membrane was obtained via hydro- polymer gel in hydrothermal conditions. This allowed to obtain
thermal treatment of geopolymers based on silica fume and met- zeolite X monoliths with relatively high compressive strength. The
akaolin (Zhang et al., 2019). ECR-1 zeolite is a synthetic zeolite authors studied the synthesis parameters such as Na2O/SiO2 molar
consisting of alternating mazzite and mordenite layers with Si/Al ratio, H2O/Na2O molar ratio, hydrothermal treatment duration and
molar ratio of 3.5e4.5. temperature, and the optimal conditions were 1.0, 70, 18 h and
It is also possible to obtain a foamed geopolymer-zeolite blocks 90  C, respectively, giving the compressive strength of 38.5 MPa.
and spheres. The first ones were obtained via hydrothermal treat- Some examples of zeolites obtained via hydrothermal treatment
ment of geopolymer foams based on fly ash and fly ash with 20 wt% with the properties of geopolymer-zeolite composites are sum-
substitution of granulated blast furnace slag (Zhang et al., 2016). marized in Table 3.
Zeolite NaeP1 was detected in the autoclaved samples. It was
observed that the addition of slag reduced the amount of zeolite

Table 3
The properties of geopolymer-zeolite composites obtained in hydrothermal conditions.

Obtained zeolites Starting Curing conditions Compressive strength BET surface area (m2/ Porosity (cm3/ Ref.
material (MPa) g) g)

NaeP1 FA (1) 60  C, 24 h 23.21 50.46 0.092 Qiu et al. (2014)


(2) HT 100  C, 24 h, 2 M
NaOH
Faujasite FA (1) RT, 12 h 5.14 174.35 0.14 (Liu et al., 2016a,
(2) 80  C, 12 h 2016b)
(3) HT 70  C, 24 h, 1 M NaOH
NaeP1 FA, BFS (1) 80  C, 24 h 16.57 114.16 0.27 Lee et al. (2016a)
(2) HT 90  C, 24 h
Faujasite, NaeP1 Biomass FA, MK (1) RT, 24 h 9.78 56.35 e De Rossi et al. (2018)
(2) HT 60  C, 28 d
ZSM-20 MK (1) 40  C, 3 d e 78.52 0.23 Tang et al. (2016)
(2) HT 140  C, 10 h, water
Zeolite X MK (1) 60  C, 24 h 19.6 63.31 0.229 Zhang et al. (2014)
(2) HT 90  C, 15 h, water
ECR-1 MK, FS (1) 50  C, 24 h e 84.26 0.019 Zhang et al. (2019)
(2) HT 180  C, 24 h, water
NaeP1 MK (1) 60  C, 3 d e 42.08 e Tang et al. (2015)
(2) HT 140  C, 10 h
NaeP1, chabazite and FA, BFS (1) HT 100  C, 48 h 15.3 76.6 0.24 Khalid et al. (2018)
analcime
Zeolite X MK (1) oven, 30 min 38.46 482 0.25 Wang et al. (2018)
(2) HT 90  C, 18 h
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 569

2.4. Mechanism of zeolites formation in geopolymers

Geopolymers structure is very similar to zeolite ones and thus


both materials are formed in a similar manner (Kro l et al., 2018; van
Jaarsveld et al., 1997). Most authors agree that it involves a three
step model of dissolution, orientation and hardening (Provis et al.,
2005; van Jaarsveld et al., 1997). Some authors claim (Pacheco-
Torgal et al., 2008) that the exact reaction mechanism of alkali-
activated binders is not yet quite understood, although it depends
on both the type of starting materials and alkaline activator.
Generally, it can be stated that CeSeH phase is the main reaction
product in the CaOeSiO2 system, while zeolite like polymers are
formed in the Al2O3eSiO2 system. Although the mechanism of
zeolite formation in geopolymers is not yet fully understood, some
models have been presented so far.
Crucial aspect in the crystallization of zeolites is water excess.
Brough et al. (1994) detected high amounts of tobermorite,
hydroxysodalite and zeolite NaeP1 in the Na2OeAl2O3eSiO2eH2O
and Na2OeCaOeAl2O3eSiO2eH2O systems characterized by high Fig. 6. Model of zeolite A formation mechanism from geopolymer spheres. Repro-
water/binder ratio. Rodríguez et al. (2013b) stated that the trans- duced with permission from Ref. (Liu et al., 2017b). Copyright 2017 John Wiley and
formation of amorphous aluminosilicate gel into zeolites is also Sons.
determined by the dissolution rate of a raw material. This is
controlled by at least two processes: (1) the breakage of surface
bonds breakage due to the solvent action, and the formation of the last stage missed due to slow dissolution and condensation
soluble species that leave the surface of the dissolving solid and (2) reactions when the hardening take place.
the reaction of the species on/with the surface of the dissolving Other mechanism of zeolite formation in geopolymeric spheres
solid (Rodríguez et al., 2013b). The occurrence of solid phase has been developed as well (Liu et al., 2017b; Pacheco-Torgal et al.,
transformation mechanism presumes the development of a poly- 2008). During the formation of zeolites, at first a stable ordered
merized hydrogel between the silicate and aluminate phases nanocrystal nucleus is formed in the geopolymer gel. This geo-
following the crystallization condition, including the rearrange- polymer is composed of N-A-S-H gel, dissolved aluminosilicate
ment of the aluminosilicate skeleton into a crystalline framework of particulates, free sodium ions, voids, and water. In high alkalinity
zeolites (Yan et al., 2012). Provis and van Deventer (Provis and van regions e where free sodium ions fill in the voids in significant
Deventer, 2007a; Provis and van Deventer, 2007b) confirmed that amounts e the aluminosilicate gel particles dissolve and generate
the initial geopolymer gel is transformed over time into a more [SiO4] and [AlO4] tetrahedra. They induce polymerization and
ordered phase. rearrange on the surface of the particles when these two types of
As mentioned, the geopolymerization-crystallization mechanism tetrahedra are sufficiently concentrated close to the theoretical
of zeolite formation involves three steps (Duan et al., 2015; van composition of a zeolite forming the ordered nanocrystalline core.
Jaarsveld et al., 1997; Xu and Van Deventer, 2003): With an increase in SieO and AleO polymers, the nanocrystals
gradually grow into zeolites. This depends mainly on the transport
1. The dissolution-hydrolysis: the release of silicate and aluminate of water molecules. They tend to diffuse to the surface of geo-
monomers under an alkaline attack, the hydrolysis firstly occurs polymer spheres. The formation of a nanocrystalline core occurs
on shells of solid particles. This is essential for the solid particles more likely in the surface of the spheres followed by gradual inside
conversion into a geopolymeric gel. growth. Fig. 6 shows the model of zeolite A formation mechanism
2. The hydrolysis-polycondensation: Then dissolved species cross- from geopolymer spheres.
link to form oligomers, which then form N-A-S-H gel. This stage
includes gel generation, and simultaneous setting and hard- 3. Resistance of zeolites in the geopolymeric matrices
ening due to polycondensation and the formation of the 3D
aluminosilicate network. The resulting geopolymer gel has a Geopolymers exhibit good thermal and fire resistance up to
zeolite-like structure. 1300  C (Provis and Van Deventer, 2009), sulfate resistance
3. The polymerization-crystallization: with the polymerization (Bakharev, 2005b; Fernandez-Jimenez et al., 2007; Grutzeck et al.,
progress, the three dimensional aluminosilicate networks keep 2002) and acid resistance (Roy, 1999). For comparison conventional
short-range ordering. During the geopolymerization water acts concrete shows some disadvantages such as low chemical resis-
as a reagent and also as a reaction byproduct e the dissolution of tance to acids and salts, low thermal and fire resistance, especially
substrates uses water, but the polymerization releases water (Xu at temperatures over 500  C. These unique properties of geo-
and Van Deventer, 2003). polymers owe to structural similarity to zeolites. In addition, the
presence of zeolite phases in the geopolymeric matrix gives this
Fernandez-Jime
nez et al. (2005) confirmed that under the in- type of composites catalytic, sorption or molecular sieve properties.
fluence of the alkaline solution on the fly ash as raw material, a Therefore, the stability of zeolite phases is important from the point
dissolution process occurs. Then, the molecules condense into a of view of the potential use of such materials in catalysis or envi-
geopolymer gel. The alkaline attack opens the fly ash spheres, ronmental protection.
exposing small spheres inside them, which will be also dissolved. Zeolite phases have limited resistance to acids. Bakharev
With the formation of reaction products inside and outside the (2005c) observed the complete disappearance of zeolitic phases
sphere the amorphous phase is completely dissolved. The same (zeolite NaeP1, hydroxysodalite) in fly ash-based geopolymers
research team stated (Palomo et al., 2008) that the geo- exposed to 5% acetic acid solution. The similar situation was
polymerization process may be considered as a zeolitization with observed in the case of exposure to 5% sulfuric acid solution, but
570 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

only for the geopolymer activated with NaOH. When sodium hy-
droxide was used together with potassium hydroxide as an acti-
vator, zeolite NaeP1 did not disappeared. Sturm et al. (2018)
observed the disappearance of zeolite A after the exposure of
microsilica-based geopolymer to sulfuric acid. Dissolution of sili-
cates in general includes: (1) leaching of non-framework cations,
i.e. the substitution of Naþ by H3Oþ in the case of zeolite A,
hydroxysodalite and geopolymeric gel; (2) removal of Al from the
framework by hydrolysis of SieOeAl bonds; (3) hydrolysis of
SieOeSi bonds, if such bonds are present in the structure. For
zeolite A, step (2) causes the complete destruction of the frame-
work, since each Si is bonded to four Al via SieOeAl bonds (only
Q4(4Al) silicate species occur in its structure). Gordon et al. (2014)
investigated the resistance of geopolymers exposed to carbon
capture solvents (monoethanolamine and potassium carbonate). In
the case of geopolymers based on metakaolin and granulated blast
furnace slag faujasite was detected in all the samples changing in
SiO2/Al2O3 ratio (3.0, 3.3 and 3.6), while Na-chabazite only in the
sample with the highest ratio. It was stated that the coexistence of
these zeolitic phases is possible due to the high Si/Al ratio, presence
of alkali-rich pore solution, and reduced extent of geo-
polymerization. Zeolites were not detected in samples exposed to
solvents, possibly due to the dilution (removal) of the alkaline pore
solution, which stopped zeolite growth reaction. The inhibition of
zeolite crystallization was also observed at the lowest Na2O content
(limited precursor dissolution) as well as reduced water content Fig. 7. The percentage of zeolite V (:) and faujasite (C) in NaK-geopolymer ther-
mally treated up to 1000  C. reproduced with permission from Ref. (Duxson et al.,
(steric restrictions). Brylewska et al. (2018) detected zeolite A and
2007). Copyright 2007 Elsevier.
hydroxysodalite in metakaolin-based geopolymers. The geo-
polymer sample was subjected to desilication (removal of Si from
the network with the use of NaOH) and dealumination (removal of spectra of geopolymers with and without zeolite A in the matrix are
Al with EDTA solution). The dealumination caused the destruction shown in Fig. 8b. To the vibrations of D6R rings in the framework of
of zeolite A, while hydroxysodalite remained unchanged, which zeolite X also bands at 880 and 750 cm1 were assigned (Kro l et al.,
suggests its higher stability. 2017a). The presence of hydroxysodalite is indicated by the pres-
Another factor affecting the durability of zeolites in the geo- ence of bands at 670 and 610 cme1 (Panagiotopoulou et al., 2015),
polymer matrices is temperature. The disappearance of hydrox- 725, 702, 663 cme1 (Kro l et al., 2017b), 739, 695, 661 cm1 (Rozek
_
ysodalite and zeolite NaeP1 was after the exposure of fly ash-based et al., 2018), and 753, 695, 667 cme1 (Fern andez-Jimenez et al.,
geopolymers to 800 and 1000  C (Bakharev, 2006). Faujasite and 2008), so between approximately 750 and 610 cme1. The forma-
zeolite V in NaK-geopolymer based on metakaolin were observed to tion of cancrinite can be related to the band at approx. 873 cme1,
disappear after heating to 200  C as shown in Fig. 7 (Duxson et al., which corresponds to stretching vibrations of OeCeO (Ye et al.,
2007). A significant structure transformation in geopolymers based 2014), and bands between 630 and 620 cme1 and at 515 cme1
on residue from the chlorosilane production was observed after were associated with vibrations of D6R structures in chabazite
heat treatment (Sturm et al., 2016). At 650  C all the peak intensities (Fernandez-Jime nez and Palomo, 2005).
of zeolite A decreased significantly, while the FWHM (full width at MAS NMR. The structure of zeolite A is formed by linking sodalite
half maximum) of the peaks increased, which can be attributed to cages through double four-membered rings, which present a signal
the structural breakdown. At 750  C a further phase transition was at around 89 ppm in 29Si NMR spectrum, which is associated with
observed. The presence of a new main phase, nepheline, was clearly the presence of Q4(4Al) units (each silicon tetrahedron is sur-
noticed, hydroxysodalite remained in the system, but no zeolite A. rounded by four aluminum tetrahedra), hence Si/Al ratio is equal to
1 (Fern andez-Jimenez et al., 2008). To hydroxysodalite, signals
4. Methods used to detect zeolites at 84 (Ferna ndez-Jimenez et al., 2008), 85 (Buchwald et al.,
2011; Garcia-Lodeiro et al., 2016) and 89 ppm (Granizo et al.,
XRD. The characteristic geopolymer “amorphous hump” (at 2014) were assigned. The FAU structure is formed by linking so-
approximately 28 2q) can be, at least in part, assigned to zeolite dalite cages through double six-membered rings. Faujasite can
nanocrystals less than 10 nm in size, which coexist with an amor- display up to five peaks, at 84, 88, 93, 98, and 103 ± 1 ppm,
phous gel phase (Provis et al., 2005). This crystallinity is below the which can be assigned to Si atoms surrounded by 4, 3, 2, 1, and
detection limit of XRD. When the conditions are favorable nano- 0 atoms of Al, respectively (Ferna ndez-Jime nez et al., 2008). In
crystals may transform into zeolites. The examples of diffraction other studies, to FAU structure peaks at 89 ppm (Zibouche et al.,
pattern corresponding to zeolite A is presented in Fig. 8a. 2009) and 84 (Buchwald et al., 2011) were assigned as well
FT-IR. The vibrations in the range 800e500 cm1 are believed to as 85, 89, 94, 99, and 103 ppm (Takeda et al., 2013). The
give structural information about the secondary building units peaks observed at the chemical shifts 87, 92, 97, 102,
(SBU) in zeolites (Buchwald et al., 2011). A band at around 560- and 107 ppm were attributed to zeolite NaeP1 (Takeda et al.,
550 cme1 is often assigned to zeolite A (Buchwald et al., 2011; 2013). The chlorosilane residue-based geopolymer-zeolite com-
Fernandez-Jime nez et al., 2008; Sturm et al., 2016; Zhang et al., posites were investigated by 1H, 23Na and 29Si MAS NMR to
2012; Ozer and Soyer-Uzun, 2015) since it corresponds to the vi- distinguish gel, zeolite phases and water species (Greiser et al.,
brations of double rings (D6R and D4R), also in faujasite (Buchwald 2017). The largest fraction of pore water was revealed in the a-
ndez-Jime
et al., 2011; Ferna nez et al., 2008; Liu et al., 2016a). The
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 571

Fig. 8. Geopolymers with and without zeolite A formed: a) XRD patterns (A e zeolite A, S e hydroxysodalite, Q e quartz, C e calcite); b) FT-IR spectra (arrow e band corresponding
to vibrations in zeolite A).

zeolites in a form of granules or powders require porous supports


for their industrial application (Khalid et al., 2018) or they should be
transformed into specific macroshapes by pressurization, sintering,
or spray coating, which increases the complexity and cost of
manufacturing. The so-shaped materials can easily expand and
break in the acid and alkaline systems (e.g. in ventilation filtration
processes) and the generated residue can block the channels,
causing the decrease in the adsorption properties (Wang et al.,
2018). Moreover, powder adsorbents cannot be easily recovered
after use, thus a solid bulk-type adsorbents, such as geopolymer-
zeolite composites, can be a better alternative.
For instance, geopolymer-zeolite composites were utilized in
the promising low temperature capture of CO2 (Minelli et al., 2018).
Although the selectivity and capacity properties were not beyond
the state of the art for the capture of CO2 by monolithic adsorbents,
the combination of Na13X zeolite filler with a geopolymer matrix
can enhance the properties of both phases. The self-supporting
porous adsorbent monoliths with sufficient strength (3 MPa) can
overcome the drawbacks of conventional granules or packed beds
Fig. 9. 29Si MAS NMR spectra of geopolymers (84 ppm assigned to faujasite and 85 sorption systems. The geopolymer-supported zeolite X has been
to hydroxysodalite; B5 spectrum typical for amorphous gel). Reproduced with
also proposed as a desiccant for desiccant wheel in air conditioning
permission from Ref. (Buchwald et al., 2011). Copyright 2011 Elsevier.
applications (Wang et al., 2018). Geopolymer with the addition of
5 A zeolite, in turn, exhibited good formaldehyde-purifying prop-
cages of the zeolite A and in the geopolymeric gel. Moreover, water erty (Huang et al., 2012).
exists in the b-cages of the zeolites and is adsorbed at sodium ions. A proposed application of the geopolymer-zeolite composites is
Fig. 9 shows 29Si MAS NMR spectra of geopolymers. a large-scale wastewater treatment. Bulk-type adsorbents can be a
SEM. Zeolitic phases are often not detected by SEM due to their substitute for traditional zeolite powder, which allows to avoid

occurrence in form of small crystals (Alvarez-Ayuso et al., 2008) or secondary powder pollution (Liu et al., 2016a). The geopolymer-
they present far from the perfect shape of pure zeolite (Zhang et al., faujasite monolith was proposed for the purification of water and
2012). However, as can be seen in Fig. 10, when the conditions of as zeolite sieves for the adsorption of heavy metals (Lee et al.,
growth are favorable, zeolites can be observed clearly. Faujasite 2016a). The geopolymer-supported zeolites were applied as sor-
exhibits octahedral crystal morphology (Fig. 10a), zeolite A forms as bents with high Pb2þ adsorption capacity (37.9 mg/g), higher than
cubic crystals (Fig. 10b), hydroxysodalite as lamellar wool ball-like that of other bulk-type adsorbents. It should be noted that the
particles (Fig. 10c), crystals of zeolite NaeP1 has diamond-like microstructure of a sample (pore size and volume, surface area)
morphology, and analcime has icositetrahedral crystals. affects the adsorption capacity e heavy metal cations have to
diffuse through the geopolymeric matrix. Adsorption is not realized
only by zeolites (ion-exchange or selective sieving), but also N-A-S-
5. Applications of geopolymer-zeolite composites H gel contributes in adsorption, although with lower capacity
(Khalid et al., 2018).
Geopolymer-zeolite hybrid materials are used as sorbents, both Liu et al. (2016b) prepared faujasite blocks via hydrothermal
in gaseous and aqueous systems. Zeolites are efficient adsorbents treatment of fly ash-based geopolymer. Geopolymer-zeolite block
due to their high surface area and cation exchange capacity (CEC). had a higher Pb2þ removal efficiency than geopolymer without
They are usually synthesized under hydrothermal conditions by zeolite, i.e. 143.3 and 118.6 mg/g, respectively (Fig. 11). Zhang et al.
alkali-activation of silica and alumina source materials. However,
572 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

Fig. 11. Pb2þ adsorption performance of fly ash (FA), fly ash-based geopolymer (GEO)
and geopolymer-supported faujasite (FAU). Reproduced with permission from Ref. (Liu
et al., 2016b). Copyright 2016 Elsevier.

(2016) treated hydrothermally geopolymer foams based on fly ash


or fly ash with blast furnace slag (20 wt% substitution). Zeolite
NaeP1 was identified in the autoclaved samples. Cation exchange
capacity (CEC) values of geopolymer-zeolite foams were measured.
CEC was substantially higher for the sample based on fly ash only
(54 vs 24 meq/100 g for the sample with slag), and was acceptable
for industrial wastewater treatment. Tang et al. (2015), in turn, used
hydrothermal treatment to obtain zeolite NaeP1 spheres from
metakaolin-based geopolymer spheres, prepared via suspension-
solidification method. Such geopolymer-zeolite spheres may be
utilized in softening water. The porous geopolymer-zeolite A
spheres were obtained in non-hydrothermal synthesis (Liu et al.,
2017b). They can be packed in column bed for continuous pro-
cesses and easily regenerated after using in the fields of adsorption
or catalysis. The utilization of geopolymer-zeolite composites in
catalysis was also suggested in the study (Brylewska et al., 2018).
Lee et al. (2017) prepared geopolymers based on fly ash and
blast furnace slag (weight ratio of 4:1). After curing at 80  C for 24 h,
they were hydrothermally treated at 90  C for 24 h. Peaks corre-
sponding to zeolite NaeP1, faujasite, chabazite and hydroxysodalite
were observed at the diffraction pattern of the final product. The
obtained mesoporous geopolymer-zeolite composite was used as
Csþ adsorbent. In the optimal conditions (a solution pH of 4, a
contact time of 24 h) the removal of Csþ was more than 90%
(adsorption capacity 15.24 mg/g). Haddad et al. (2017) applied
long-term curing of geopolymers at low temperature (40  C for 1
week, 1 month and 3 months). Moreover, solutions of CsOH in
different proportions with NaOH were used as activators in order to
assess the geopolymer ability of cesium immobilization. The ob-
tained geopolymers with 50 and 7% Cs replacement were
composed of an amorphous phase as well as Cs-bearing zeolite
being an analogue of zeolite F (framework type EDI). Zeolite A and
zeolite X were present in 7%-sample as well and were the only
zeolitic phases in the samples with 1% and 0% Cs replacement.
Leaching experiments showed that Cs is efficiently bound by zeolite
F and geopolymers can be utilized as waste forms for 137Cs. Geo-
polymers with Cs-bearing zeolites have been also described in the
studies (Bell et al., 2008; Ofer-Rozovsky et al., 2016; Shiota et al.,
Fig. 10. SEM images of zeolites formed in geopolymer matrices: a) faujasite (repro- 2017a).
duced with permission from Ref. (Liu et al., 2016b). Copyright 2016 Elsevier); b) zeolite An important and promising field of geopolymer-zeolite com-
A; c) hydroxysodalite.
posites is their utilization as self-supporting membranes. Nowadays,
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 573

membrane separation processes have found wide applications in


various fields. This separation is induced by the differences in
chemical and physical properties of the membrane and the perme-
able component. The permeation driving force can be concentration,
pressure, temperature or electrical potential gradient (Azarshab
et al., 2016). Zeolite membranes, which have been intensively
studied in recent decades, exhibit good separation properties in gas
permeation, dehydration and pervaporation processes (Xu et al.,
2017) and several advantages: (1) they do not swell like polymeric
membranes; (2) in strong alkaline or acidic systems they are more
chemically stable than polymeric membranes; (3) zeolites are more
stable at relatively high temperatures (Yan et al., 2012). The main
disadvantage of zeolite membranes is a costly and non-reproducible
synthesis process (Azarshab et al., 2016) and they are generally more
brittle than polymers (Yan et al., 2012).
Pervaporation, one of the membrane processes, is an energy-
efficient and effective technique, introduced by Kobber in 1917.
The mechanism of separation in a pervaporation process is
solution-diffusion. It is a method for conducting separations of Fig. 12. Schematic of the self-supporting zeolite membrane with non-zeolitic pores.
mixtures of liquids by partial evaporation through a membrane Reproduced with permission from Ref. (He et al., 2013). Copyright 2013 Elsevier.
(feed side). Usually, a carrier gas is used or a vacuum is created to
reduce a partial pressure in permeate side. This process has gained
increasing interest for several industrial applications and is used for of the geopolymer membrane was nearly zero, while in the case of
the removal of water from organics (dehydration of hydrocarbons). the geopolymer-zeolite membrane the water flux was over
As a result a high purity product can be obtained, e.g. ethanol, 2 kg me2$he1 (20  C). Most of the geopolymer gel converted into
isopropyl alcohol and ethylene glycol (Azarshab et al., 2016). Since the zeolite phase after the hydrothermal process, while the rest of
zeolites contain uniform, molecular-sized pores, zeolite mem- the gel continued to provide strength for the membrane. The pre-
branes are suitable for pervaporation (Yan et al., 2012). Zeolite liminary separation results showed that ethanol solutions can be
membranes can be formed as zeolite layers on microporous inor- significantly dehydrated by pervaporation. Also, the faujasite
ganic supports (e.g. silica, alumina or stainless steel) (He et al., membrane was found to effectively separate alcohol from water. A
2013). These layers can be obtained by means of methods like self-supporting faujasite membrane was obtained via in situ hy-
sol-gel, microwave and hydrothermal synthesis, which induce the drothermal treatment of a geopolymer and exhibited high strength,
growth of zeolite crystals. In the hydrothermal method supports good stability and efficient separation capability (water flux of
seeded with zeolite precursor are kept in an autoclave at specified 3.16 kg me2$he1 at 80  C) (Zhang et al., 2014). Yugawaralite zeolite
temperatures (Azarshab et al., 2016). It is a highly efficient method membrane was successfully used in the pervaporation separation
but defect-free membranes are hard to be obtained (Azarshab et al., of ethanol/water mixture as well due to its low SiO2/Al2O3 ratio that
2016; Xu et al., 2017). Because zeolite A membranes are highly makes it hydrophilic (Azarshab et al., 2016). Also, ECR-1 zeolite
hydrophilic and selective, with pore size of 0.4 nm, they are suitable membrane was successfully used for seawater desalination (Zhang
for the removal of water from organic solutions by pervaporation et al., 2019). The porous geopolymer-zeolite composites based on
(Yan et al., 2012; Xu et al., 2017). Such membranes must have suf- metakaolin and biomass fly ash were proposed not only for
ficient mechanical strength without macroscopic defects, so self- membranes, but also for the adsorption of volatile organic com-
supported membranes in the form of geopolymer-zeolite com- pounds and toxic metal compounds as well as CO2 capture (De
posites are promising for this application, and can be considered as Rossi et al., 2018).
a substitute for self-supported zeolite membranes that are highly The application of geopolymer-zeolite composites as construc-
fragile and lack strength. Self-supporting membranes are tion and building materials is not very popular, mostly due to un-
composed of many zeolite layers (L1, L2 … …Ln in the vertical di- favorable mechanical properties of such materials. They may be,
rection, R1, R2 … ….Rn in the horizontal direction), including non- however, used for heat insulating coatings, which is related to the
zeolitic pores (He et al., 2013), as shown in Fig. 12. The number of release of zeolitic water during heating. As such, zeolite-
non-zeolitic pores is limited, and the distribution of these pores is geopolymer hybrid materials can be also used as building mate-
random, so the effect of non-zeolitic pore defects in the rials with the property of humidity control (Takeda et al., 2013).
geopolymer-supporting zeolite membrane could be eliminated Such materials can be also used as precursors for nepheline-type
when the thickness of the zeolite membrane is of a certain size. ceramics manufacturing (Sturm et al., 2016). The alkali-activated
Xu et al. (2017) prepared zeolite A membrane by geopolymer- hydroceramics based on slag-metakaolin-fly ash systems were
gel-thermal-conversion (GGTC) without the need of hydrothermal used for waste immobilization as well and had a satisfying aqueous
treatment. The geopolymeric gel was obtained by the alkali- stability and a low leachability (Wang et al., 2016).
activation of chemosynthesized Al2O3eSiO2 powders and it was
coated on stainless steel supports in the process of dip-coating and 6. Influence of zeolites on the properties of geopolymer-
then cured at 60  C for 6 h. The geopolymer-supported zeolite A zeolite composites
membrane had a pervaporation effect on the desalination of NaCl
solution if it has a suitable thickness (9.4 mm with ion rejection of Table 4 presents the comparison of properties between geo-
99.5%) and high compressive strength (57 MPa). It can be therefore polymers without zeolites and geopolymer-zeolite hybrid mate-
used as a separation membrane for seawater desalination (water rials. As can be seen the presence of zeolites in some cases enhances
flux of 3.86 kg me2$he1 at 90  C) (He et al., 2013). Yan et al. (2012) the compressive strength, but in some cases decreases it. The
compared the geopolymer membrane and the geopolymer- transformation of amorphous gel into zeolites causes the reorga-
supported zeolite A membrane. It was shown that the water flux nization of local structure affecting the mechanical properties. It
574 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

Table 4
Geopolymer without zeolites vs. geopolymer-zeolite composites e exemplary properties.

Geopolymer without zeolites Geopolymer-zeolite composite Zeolite Ref.

Compressive strength (MPa)


18.70 5.14 faujasite Liu et al. (2016a)
7.5 3.0 zeolite X, zeolite A Papa et al. (2018)
10.05 9.78 faujasite, NaeP1 De Rossi et al. (2018)
21.50 23.21 NaeP1 Qiu et al. (2014)
16.3 19.6 zeolite X Zhang et al. (2014)
21.66 32.46 zeolite X Wang et al. (2018)
49.1 57.6 zeolite A He et al. (2013)
BET surface area (m2/g)
1.07 174.35 faujasite Liu et al. (2016a)
13 211 zeolite X, zeolite A Papa et al. (2018)
40.69 56.35 faujasite, NaeP1 De Rossi et al. (2018)
0.26 50.46 NaeP1 Qiu et al. (2014)
4.18 78.52 ZSM-20 Tang et al. (2016)
0.43 63.31 zeolite X Zhang et al. (2014)
1.84 84.26 ECR-1 Zhang et al. (2019)
4.21 42.08 NaeP1 Tang et al. (2015)
0.4 482 zeolite X Wang et al. (2018)
20.48 174.35 faujasite Liu et al. (2016b)
6.57 97.81 zeolite A He et al. (2013)
Bulk density (g/cm3)
1.15 0.99 zeolite X, zeolite A Papa et al. (2018)
1.09 1.16 faujasite, NaeP1 De Rossi et al. (2018)
Apparent porosity (%)
38 47 zeolite X, zeolite A Papa et al. (2018)
44.36 50.38 faujasite, NaeP1 De Rossi et al. (2018)
Pore volume (cm3/g)
0.0024 0.14 faujasite Liu et al. (2016a)
0.302 0.470 zeolite X, zeolite A Papa et al. (2018)
0.0006 0.092 NaeP1 Qiu et al. (2014)
0.012 0.23 ZSM-20 Tang et al. (2016)
0.000386 0.229 zeolite X Zhang et al. (2014)
0.0011 0.019 ECR-1 Zhang et al. (2019)
0.001 0.25 zeolite X Wang et al. (2018)
0.070 0.14 faujasite Liu et al. (2016b)
Water absorption (%)
42.20 43.53 faujasite, NaeP1 De Rossi et al. (2018)

can be assumed that there is a certain limit of zeolite content that micromorphological features are also an important factor (Takeda
can be bear by a geopolymeric matrix, after which it causes its et al., 2013).

weakening (Alvarez-Ayuso et al., 2008). This weakening may be the To some point, the presence of zeolites in the obtained materials
result of bonding strength. Moreover, zeolite crystalline phases may enhanced the compressive strength. However, a higher content of
act as a defect within the geopolymer matrix (Bortnovsky et al., zeolites, for instance in samples treated hydrothermally for 7 days,
2008). The compressive strength of geopolymers decreased also decreased the compressive strength (Kolousek et al., 2007). Intense
with increasing zeolite crystallite size (Fig. 13), suggesting that zeolite formation reduces compressive strength due to their
microporous-crystalline structure based on 3D-cage system when
compared to the amorphous structure of geopolymer based on 3D-
network of silicate and aluminate tetrahedra (Istuque et al., 2016).
The relationship between amorphous product content and
compressive strength is presented in Fig. 14. However, in the study

Fig. 14. Comparison of amorphous reaction product content with compressive


Fig. 13. Relationship between the compressive strength and the crystallite size of strength. Reproduced with permission from Ref. (Bhagath Singh and Subramaniam,
zeolite in final materials (Takeda et al., 2013). 2016). Copyright 2016 Elsevier.
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 575

hybrid bulk materials take advantage of both constituents: high


surface area, pore volume and adsorption capacity from zeolites,
and strength and support from geopolymer matrix. They can be
therefore used as bulk-type adsorbents, both in gaseous and
aqueous systems, and self-supporting membranes. There is a wide
range of research methods to study both the structure and prop-
erties of such materials. Since industrial by-products (ash, slag) can
be used for their synthesis, geopolymer-zeolite hybrid materials
exhibit some production and sustainability benefits: low cost and
efficient waste management.
This article reviews research papers published until the end of
2018. It presents an overview of the geopolymer-zeolite composites
synthesis approaches. They can be obtained in three ways: (1)
utilization of zeolites as raw materials for geopolymer synthesis;
(2) direct formation of zeolites parallel to geopolymerization; (3)
Fig. 15. Relationship between zeolite NaeP1 content in geopolymers and BET surface hydrothermal treatment of geopolymers. The main factors affecting
area of composites. Reproduced with permission from Ref. (Lee et al., 2016a). Copyright zeolite formation in geopolymeric matrix are: Si/Al ratio, the type
2016 Elsevier.
and content of an activator, curing conditions, the presence of ad-
ditives. Since all these issues have been discussed in this review, it
(Takeda et al., 2013) the effect of zeolites content on the can be considered as a practical guide to geopolymer-zeolite
compressive strength of a composite was not very clear. The composites manufacturing. Some issues, however, need further
compressive strength of geopolymer can be increased because of research. The application of these hybrid materials in the field of
the increased crystallization contacts between zeolite and amor- catalysis and gas separation is very promising, but still barely

phous phase (Alvarez-Ayuso et al., 2008). investigated.
The formation of zeolites contributed to a significant increase in
BET surface area of geopolymer-zeolite composites compared to Declarations of interest
geopolymer without zeolites in every case presented in Table 4. A
particular example of the effect zeolite NaeP1 content on BET None.
surface area is shown in Fig. 15. Zeolitic crystals accumulate to form
a large number of micro-pores and meso-pores, causing an increase Acknowledgement
in BET surface area (Tang et al., 2016). For comparison BET surface
area of commercial zeolite Na-13X is 791 m2/g (Papa et al., 2018; The authors acknowledges financial support from the National
Minelli et al., 2018), and of commercial 13X molecular sieve Science Centre in Poland, Grant No. 2015/17/B/ST8/01200.
spheres is 545 m2/g with pore volume of 0.26 cm3/g (Wang et al.,
2018). The presence of zeolites in a geopolymeric matrix affects References
bulk density and apparent porosity as well. Generally, the values of
Alshaaer, M., El-Eswed, B., Yousef, R.I., Khalili, F., Rahier, H., 2016. Development of
porosity (%) and pore volume (cm3/g) increases in such composites functional geopolymers for water purification, and construction purposes.
when compared to geopolymers without zeolites. J. Saudi Chem. Soc. 20 https://doi.org/10.1016/j.jscs.2012.09.012. S85eS92.

Alvarez-Ayuso, E., Querol, X., Plana, F., Alastuey, A., Moreno, N., Izquierdo, M.,
The weak adhesion of the geopolymeric coatings to the metal
Font, O., Moreno, T., Diez, S., Va zquez, E., Barra, M., 2008. Environmental,
substrates was related to the formation of zeolite A in metakaolin- physical and structural characterisation of geopolymer matrixes synthesised
based geopolymer (Si/Al ratio of 1) (Temuujin et al., 2009) and from coal (co-)combustion fly ashes. J. Hazard Mater. 154, 175e183. https://
chabazite in fly ash-based geopolymer (Si/Al ratio of 1, too) doi.org/10.1016/j.jhazmat.2007.10.008.
Alzeer, M.I.M., MacKenzie, K.J.D., 2018. Synthesis and catalytic properties of new
(Temuujin et al., 2010). The formation of zeolite (faujasite) was also sustainable aluminosilicate heterogeneous catalysts derived from fly ash. ACS
found to be responsible for the increase in shrinkage values of Sustain. Chem. Eng. 6, 5273e5282. https://doi.org/10.1021/
geopolymers (Khater and Abd El Gawaad, 2016). This effect was acssuschemeng.7b04923.
Andrejkovi , S., Sudagar, A., Rocha, J., Patinha, C., Hajjaji, W., Da Silva, E.F.,
cova
significantly lowered by the addition of carbon nanotubes. Zeolites Velosa, A., Rocha, F., 2016. The effect of natural zeolite on microstructure, me-
were also partly responsible for the degradation of metakaolin- chanical and heavy metals adsorption properties of metakaolin based geo-
based geopolymeric foams because they are metastable compo- polymers. Appl. Clay Sci. 126, 141e152. https://doi.org/10.1016/
j.clay.2016.03.009.
nents (Delair et al., 2012).
Auqui, N.U., Baykara, H., Rigail, A., Cornejo, M.H., Villalba, J.L., 2017. An investigation
The properties of geopolymer-zeolite composites depend on the of the effect of migratory type corrosion inhibitor on mechanical properties of
quantities of zeolites that formed in a matrix. These can vary zeolite-based novel geopolymers. J. Mol. Struct. 1146, 814e820. https://doi.org/
significantly and are determined by the reaction parameters. For 10.1016/j.molstruc.2017.06.066.
Azarshab, M., Mohammadi, F., Maghsoodloorad, H., Mohammadi, T., 2016. Ceramic
example fly ash/slag-based geopolymer treated hydrothermally membrane synthesis based on alkali activated blast furnace slag for separation
contained of 11.6% zeolite NaeP1 and 7.1% sodalite (Lee et al., of water from ethanol. Ceram. Int. 42, 15568e15574. https://doi.org/10.1016/
2016a) or 13% zeolite NaeP1, 1% chabazite and 1% analcime j.ceramint.2016.07.005.
Azevedo, A.G. de S., Strecker, K., 2017. Brazilian fly ash based inorganic polymers
(Khalid et al., 2018), metakaolin-based geopolymers 80.6% zeolite production using different alkali activator solutions. Ceram. Int. 43, 9012e9018.
NaeP or 35.3% zeolite X (Takeda et al., 2013), and fly ash-based https://doi.org/10.1016/j.ceramint.2017.04.044.
geopolymer 30% Na-chabazite, 1.2% cancrinite and 1.8% hydrox- Baerlocher, C., McCusker, L.B.. Database of Zeolite Structures. n.d. http://www.iza-
structure.org/databases/ (accessed March 1, 2018).
ysodalite (Criado et al., 2007) (by weight). Bakharev, T., 2005. Geopolymeric materials prepared using Class F fly ash and
elevated temperature curing. Cement Concr. Res. 35, 1224e1232. https://
doi.org/10.1016/j.cemconres.2004.06.031.
7. Summary Bakharev, T., 2005. Durability of geopolymer materials in sodium and magnesium
sulfate solutions. Cement Concr. Res. 35, 1233e1246. https://doi.org/10.1016/
j.cemconres.2004.09.002.
The potential of geopolymer conversion to zeolite can be Bakharev, T., 2005. Resistance of geopolymer materials to acid attack. Cement
exploited for the synthesis of geopolymer-zeolite composites. Such Concr. Res. 35, 658e670. https://doi.org/10.1016/j.cemconres.2004.06.005.
576 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

Bakharev, T., 2006. Thermal behaviour of geopolymers prepared using class F fly ash Duxson, P., Lukey, G.C., van Deventer, J.S.J., 2007. The thermal evolution of meta-
and elevated temperature curing. Cement Concr. Res. 36, 1134e1147. https:// kaolin geopolymers: Part 2 - phase stability and structural development. J. Non-
doi.org/10.1016/j.cemconres.2006.03.022. Cryst. Solids 353, 2186e2200. https://doi.org/10.1016/j.jnoncrysol.2007.02.050.
Baykara, H., Cornejo, M.H., Murillo, R., Gavilanes, A., Paredes, C., Elsen, J., 2017. Eiamwijit, M., Pachana, K., Kaewpirom, S., Rattanasak, U., Chindaprasirt, P., 2015.
Preparation, characterization and reaction kinetics of green cement: Ecuadorian Comparative study on morphology of ground sub-bituminus FBC fly ash geo-
natural mordenite-based geopolymers. Mater. Struct. 50, 1e12. https://doi.org/ polymeric material. Adv. Powder Technol. 26, 1053e1057. https://doi.org/
10.1617/s11527-017-1057-z. 10.1016/j.apt.2015.04.013.
Bell, J.L., Sarin, P., Provis, J.L., Haggerty, R.P., Driemeyer, P.E., Chupas, P.J., Van El-Eswed, B.I., Yousef, R.I., Alshaaer, M., Hamadneh, I., Al-Gharabli, S.I., Khalili, F.,
Deventer, J.S.J., Kriven, W.M., 2008. Atomic structure of a cesium aluminosilicate 2015. Stabilization/solidification of heavy metals in kaolin/zeolite based geo-
geopolymer: a pair distribution function study. Chem. Mater. 20, 4768e4776. polymers. Int. J. Miner. Process. 137, 34e42. https://doi.org/10.1016/
https://doi.org/10.1021/cm703369s. j.minpro.2015.03.002.
Belviso, C., Giannossa, L.C., Huertas, F.J., Lettino, A., Mangone, A., Fiore, S., 2015. Elert, K., Pardo, E.S., Rodriguez-Navarro, C., 2015. Alkaline activation as an alter-
Synthesis of zeolites at low temperatures in fly ash-kaolinite mixtures. Micro- native method for the consolidation of earthen architecture. J. Cult. Herit. 16,
porous Mesoporous Mater. 212, 35e47. https://doi.org/10.1016/ 461e469. https://doi.org/10.1016/j.culher.2014.09.012.
j.micromeso.2015.03.012. Fernandez Jiminez, A.M., Lachowski, E.E., Palomo, A., Macphee, D.E., 2004. Micro-
Bernal, S.A., Provis, J.L., Rose, V., Mejía De Gutierrez, R., 2011. Evolution of binder structural characterisation of alkali-activated PFA matrices for waste immobi-
structure in sodium silicate-activated slag-metakaolin blends. Cement Concr. lisation. Cement Concr. Compos. 26, 1001e1006. https://doi.org/10.1016/
Compos. 33, 46e54. https://doi.org/10.1016/j.cemconcomp.2010.09.004. j.cemconcomp.2004.02.034.
Bhagath Singh, G.V.P., Subramaniam, K.V.L., 2016. Quantitative XRD study of Fernandez-Jime nez, A., Palomo, A., 2005. Mid-infrared spectroscopic studies of
amorphous phase in alkali activated low calcium siliceous fly ash. Constr. Build. alkali-activated fly ash structure. Microporous Mesoporous Mater. 86, 207e214.
Mater. 124, 139e147. https://doi.org/10.1016/j.conbuildmat.2016.07.081. https://doi.org/10.1016/j.micromeso.2005.05.057.
Bhagath Singh, G.V.P., Subramaniam, K.V.L., 2017. Evaluation of sodium content and Fernandez-Jime nez, A., Palomo, A., Criado, M., 2005. Microstructure development of
sodium hydroxide molarity on compressive strength of alkali activated low- alkali-activated fly ash cement: a descriptive model. Cement Concr. Res. 35,
calcium fly ash. Cement Concr. Compos. 81, 122e132. https://doi.org/10.1016/ 1204e1209. https://doi.org/10.1016/j.cemconres.2004.08.021.
j.cemconcomp.2017.05.001. Fernandez-Jime nez, A., Palomo, A., Sobrados, I., Sanz, J., 2006. The role played by the
Bobirica, C., Shim, J.H., Pyeon, J.H., Park, J.Y., 2015. Influence of waste glass on the reactive alumina content in the alkaline activation of fly ashes. Microporous
microstructure and strength of inorganic polymers. Ceram. Int. 41, Mesoporous Mater. 91, 111e119. https://doi.org/10.1016/
13638e13649. https://doi.org/10.1016/j.ceramint.2015.07.160. j.micromeso.2005.11.015.
€ke, N., Birch, G.D., Nyale, S.M., Petrik, L.F., 2015. New synthesis method for the
Bo Fernandez-Jimenez, A., García-Lodeiro, I., Palomo, A., 2007. Durability of alkali-
production of coal fly ash-based foamed geopolymers. Constr. Build. Mater. 75, activated fly ash cementitious materials. J. Mater. Sci. 42, 3055e3065. https://
189e199. https://doi.org/10.1016/j.conbuildmat.2014.07.041. doi.org/10.1007/s10853-006-0584-8.
Boonjaeng, S., Chindaprasirt, P., Pimraksa, K., 2014. Lime-calcined clay materials Fernandez-Jime nez, A., Monzo  , M., Vicent, M., Barba, A., Palomo, A., 2008. Alkaline
with alkaline activation: phase development and reaction transition zone. Appl. activation of metakaolin-fly ash mixtures: obtain of Zeoceramics and Zeoce-
Clay Sci. 95, 357e364. https://doi.org/10.1016/j.clay.2014.05.002. ments. Microporous Mesoporous Mater. 108, 41e49. https://doi.org/10.1016/
Boonserm, K., Sata, V., Pimraksa, K., Chindaprasirt, P., 2012. Improved geo- j.micromeso.2007.03.024.
polymerization of bottom ash by incorporating fly ash and using waste gypsum Fernandez-Jime nez, A., Cristelo, N., Miranda, T., Palomo, A., 2017. Sustainable alkali
as additive. Cement Concr. Compos. 34, 819e824. https://doi.org/10.1016/ activated materials: precursor and activator derived from industrial wastes.
j.cemconcomp.2012.04.001. J. Clean. Prod. 162, 1200e1209. https://doi.org/10.1016/j.jclepro.2017.06.151.
Bortnovsky, O., Dedecek, J., Tvaru zkova, Z., Sobalík, Z., Subrt, J., 2008. Metal ions as Fletcher, R.A., MacKenzie, K.J.D., Nicholson, C.L., Shimada, S., 2005. The composition
probes for characterization of geopolymer materials. J. Am. Ceram. Soc. 3057, range of aluminosilicate geopolymers. J. Eur. Ceram. Soc. 25, 1471e1477. https://
3052e3057. https://doi.org/10.1111/j.1551-2916.2008.02577.x. doi.org/10.1016/j.jeurceramsoc.2004.06.001.
Brough, A., Katz, A., Bakharev, T., Sun, G., Kirkpatrick, R., Struble, L., Young, J., 1994. Garcia-Lodeiro, I., Aparicio-Rebollo, E., Ferna ndez-Jimenez, A., Palomo, A., 2016.
Microstructural aspects of zeolite formation in alkali activated cements con- Effect of calcium on the alkaline activation of aluminosilicate glass. Ceram. Int.
taining high levels of fly ash. MRS Proc 370, 199e208. https://doi.org/10.1557/ 42, 7697e7707. https://doi.org/10.1016/j.ceramint.2016.01.184.
PROC-370-197. Ge, Y.Y., Tang, Q., Cui, X.M., He, Y., Zhang, J., 2014. Preparation of large-sized anal-
Brylewska, K., Rozek, _ P., Krol, M., Mozgawa, W., 2018. The influence of deal- cime single crystals using the Geopolymer-Gels-Conversion (GGC) method.
umination/desilication on structural properties of metakaolin-based geo- Mater. Lett. 135, 15e18. https://doi.org/10.1016/j.matlet.2014.07.122.
polymers. Ceram. Int. 44, 12853e12861. https://doi.org/10.1016/ Geetha, S., Ramamurthy, K., 2013. Properties of geopolymerised low-calcium bot-
j.ceramint.2018.04.095. tom ash aggregate cured at ambient temperature. Cement Concr. Compos. 43,
Buchwald, A., Zellmann, H.D., Kaps, C., 2011. Condensation of aluminosilicate gels- 20e30. https://doi.org/10.1016/j.cemconcomp.2013.06.007.
model system for geopolymer binders. J. Non-Cryst. Solids 357, 1376e1382. Glid, M., Sobrados, I., Ben Rhaiem, H., Sanz, J., Amara, A.B.H., 2017. Alkaline acti-
https://doi.org/10.1016/j.jnoncrysol.2010.12.036. vation of metakaolinite-silica mixtures: role of dissolved silica concentration on
Chindaprasirt, P., Paisitsrisawat, P., Rattanasak, U., 2014. Strength and resistance to the formation of geopolymers. Ceram. Int. 43, 12641e12650. https://doi.org/
sulfate and sulfuric acid of ground fluidized bed combustion fly ash-silica fume 10.1016/j.ceramint.2017.06.144.
alkali-Activated composite. Adv. Powder Technol. 25, 1087e1093. https:// Gordon, L.E., Nicolas, R.S., Provis, J.L., 2014. Chemical characterisation of metakaolin
doi.org/10.1016/j.apt.2014.02.007. and fly ash based geopolymers during exposure to solvents used in carbon
Clausi, M., Fern andez-Jime nez, A.M., Palomo, A., Tarantino, S.C., Zema, M., 2018. capture. Int. J. Greenh. Gas Control. 27, 255e266. https://doi.org/10.1016/
Reuse of waste sandstone sludge via alkali activation in matrices of fly ash and j.ijggc.2014.06.005.
metakaolin. Constr. Build. Mater. 172, 212e223. https://doi.org/10.1016/ Granizo, N., Palomo, A., Fernandez-Jime nez, A., 2014. Effect of temperature and
j.conbuildmat.2018.03.221. alkaline concentration on metakaolin leaching kinetics. Ceram. Int. 40,
Criado, M., Fern andez-Jime nez, A., de la Torre, A.G., Aranda, M.A.G., Palomo, A., 8975e8985. https://doi.org/10.1016/j.ceramint.2014.02.071.
2007. An XRD study of the effect of the SiO2/Na2O ratio on the alkali activation Greiser, S., Sturm, P., Gluth, G.J.G., Hunger, M., Ja€ger, C., 2017. Differentiation of the
of fly ash. Cement Concr. Res. 37, 671e679. https://doi.org/10.1016/ solid-state NMR signals of gel, zeolite phases and water species in geopolymer-
j.cemconres.2007.01.013. zeolite composites. Ceram. Int. 43, 2202e2208. https://doi.org/10.1016/
Criado, M., Jime nez, A.F., Palomo, A., 2010. Effect of sodium sulfate on the alkali j.ceramint.2016.11.004.
activation of fly ash. Cement Concr. Compos. 32, 589e594. https://doi.org/ Grutzeck, M.W., Puertas, F., Blanco-Varela, M.T., Granizo, M.L., Vazquez, T.,
10.1016/j.cemconcomp.2010.05.002. Palomo, A., 2002. Chemical stability of cementitious materials based on meta-
Cwirzen, A., Provis, J.L., Penttala, V., Habermehl-Cwirzen, K., 2014. The effect of kaolin. Cement Concr. Res. 29, 997e1004. https://doi.org/10.1016/s0008-
limestone on sodium hydroxide-activated metakaolin-based geopolymers. 8846(99)00074-5.
Constr. Build. Mater. 66, 53e62. https://doi.org/10.1016/ Guo, X., Shi, H., Dick, W.A., 2010. Compressive strength and microstructural char-
j.conbuildmat.2014.05.022. acteristics of class C fly ash geopolymer. Cement Concr. Compos. 32, 142e147.

Delair, S., Prud’homme, E., Peyratout, C., Smith, A., Michaud, P., Eloy, L., Joussein, E., https://doi.org/10.1016/j.cemconcomp.2009.11.003.
Rossignol, S., 2012. Durability of inorganic foam in solution: the role of alkali Guo, X., Shi, H., Chen, L., Dick, W.A., 2010. Alkali-activated complex binders from
elements in the geopolymer network. Corros. Sci. 59, 213e221. https://doi.org/ class C fly ash and Ca-containing admixtures. J. Hazard Mater. 173, 480e486.
10.1016/j.corsci.2012.03.002. https://doi.org/10.1016/j.jhazmat.2009.08.110.
Van Deventer, J.S.J., Lukey, G.C., Xu, H., 2006. Effect of curing temperature and sil- Haddad, M.A., Ofer-Rozovsky, E., Bar-Nes, G., Borojovich, E.J.C., Nikolski, A.,
icate concentration on fly-ash-based geopolymerization effect of curing tem- Mogiliansky, D., Katz, A., 2017. Formation of zeolites in metakaolin-based geo-
perature and silicate concentration on fly-ash-based geopolymerization. Ind. polymers and their potential application for Cs immobilization. J. Nucl. Mater.
Eng. Chem. Res. 45, 3559e3568. https://doi.org/10.1021/ie051251p. 493, 168e179. https://doi.org/10.1016/j.jnucmat.2017.05.046.
van Deventer, J.S.J., Provis, J.L., Duxson, P., Lukey, G.C., 2007. Reaction mechanisms in El Hafid, K., Hajjaji, M., 2015. Effects of the experimental factors on the micro-
the geopolymeric conversion of inorganic waste to useful products. J. Hazard structure and the properties of cured alkali-activated heated clay. Appl. Clay Sci.
Mater. 139, 506e513. https://doi.org/10.1016/j.jhazmat.2006.02.044. 116e117, 202e210. https://doi.org/10.1016/j.clay.2015.03.015.
Duan, J., Li, J., Lu, Z., 2015. One-step facile synthesis of bulk zeolite A through El Hafid, K., Hajjaji, M., 2018. Geopolymerization of glass- and silicate-containing
metakaolin-based geopolymer gels. J. Porous Mater. 22, 1519e1526. https:// heated clay. Constr. Build. Mater. 159, 598e609. https://doi.org/10.1016/
doi.org/10.1007/s10934-015-0034-6. j.conbuildmat.2017.11.018.
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 577

El Hafid, K., Hajjaji, M., El Hafid, H., 2017. Influence of NaOH concentration on Kro l, M., Rozek,
_ P., 2019. The effect of calcination temperature on the metakaolin
microstructure and properties of cured alkali-activated calcined clay. J. Build. structure for the synthesis of zeolites. Clay Miner. 1e7. https://doi.org/10.1180/
Eng. 11, 158e165. https://doi.org/10.1016/j.jobe.2017.04.012. clm.2018.49.
Hajimohammadi, A., van Deventer, J.S.J., 2017. Solid reactant-based geopolymers Kro l, M., Mozgawa, W., Jaste˛ rzbski, W., Barczyk, K., 2012. Application of IR spectra in
from rice hull ash and sodium aluminate. Waste Biomass Biomass 8, the studies of zeolites from D4R and D6R structural groups. Microporous
2131e2140. https://doi.org/10.1007/s12649-016-9735-6. Mesoporous Mater. 156, 181e188. https://doi.org/10.1016/
Hajimohammadi, A., Provis, J.L., Van Deventer, J.S.J., 2011. The effect of silica avail- j.micromeso.2012.02.040.
ability on the mechanism of geopolymerisation. Cement Concr. Res. 41, Kro l, M., Minkiewicz, J., Mozgawa, W., 2016. IR spectroscopy studies of zeolites in
210e216. https://doi.org/10.1016/j.cemconres.2011.02.001. geopolymeric materials derived from kaolinite. J. Mol. Struct. 1126, 200e206.
Hajimohammadi, A., Provis, J.L., van Deventer, J.S.J., 2011. Time-resolved and https://doi.org/10.1016/j.molstruc.2016.02.027.
spatially-resolved infrared spectroscopic observation of seeded nucleation Kro l, M., Brylewska, K., Knapik, A., Kornaus, K., Mozgawa, W., 2017. Conditions of
controlling geopolymer gel formation. J. Colloid Interface Sci. 357, 384e392. synthesis and structure of metakaolin-based geopolymers: application as heavy
https://doi.org/10.1016/j.jcis.2011.02.045. metal cation sorbent. Pol. J. Chem. Technol. 19 https://doi.org/10.1515/pjct-
He, Y., Cui, X., Liu, X., Wang, Y., Zhang, J., Liu, K., 2013. Preparation of self-supporting 2017-0075.
NaA zeolite membranes using geopolymers. J. Membr. Sci. 447, 66e72. https:// Kro l, M., Rozek,_ P., Mozgawa, W., 2017. Synthesis of the sodalite by geo-
doi.org/10.1016/j.memsci.2013.07.027. polymerization process using coal fly ash. Pol. J. Environ. Stud. 26, 2611e2617.
Heah, C.Y., Kamarudin, H., Mustafa Al Bakri, A.M., Bnhussain, M., Luqman, M., https://doi.org/10.15244/pjoes/70231.
Khairul Nizar, I., Ruzaidi, C.M., Liew, Y.M., 2012. Study on solids-to-liquid and Kro l, M., Rozek,
_ P., Chlebda, D., Mozgawa, W., 2018. Influence of alkali metal cations/
alkaline activator ratios on kaolin-based geopolymers. Constr. Build. Mater. 35, type of activator on the structure of alkali-activated fly ash e ATR-FTIR studies.
912e922. https://doi.org/10.1016/j.conbuildmat.2012.04.102. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 198, 33e37. https://doi.org/
Heidrich, C., Feuerborn, H., Weir, A., 2013. Coal combustion products: a global 10.1016/j.saa.2018.02.067.
perspective. In: Proc. 2013 World Coal Ash Conf., Lexington, KY. Lancellotti, I., Catauro, M., Ponzoni, C., Bollino, F., Leonelli, C., 2013. Inorganic
Hu, M., Zhu, X., Long, F., 2009. Alkali-activated fly ash-based geopolymers with polymers from alkali activation of metakaolin: effect of setting and curing on
zeolite or bentonite as additives. Cement Concr. Compos. 31, 762e768. https:// structure. J. Solid State Chem. 200, 341e348. https://doi.org/10.1016/
doi.org/10.1016/j.cemconcomp.2009.07.006. j.jssc.2013.02.003.
Hu, N., Bernsmeier, D., Grathoff, G.H., Warr, L.N., 2017. The influence of alkali acti- Lee, N.K., Khalid, H.R., Lee, H.K., 2016. Synthesis of mesoporous geopolymers con-
vator type, curing temperature and gibbsite on the geopolymerization of an taining zeolite phases by a hydrothermal treatment. Microporous Mesoporous
interstratified illite-smectite rich clay from Friedland. Appl. Clay Sci. 135, Mater. 229, 22e30. https://doi.org/10.1016/j.micromeso.2016.04.016.
386e393. https://doi.org/10.1016/j.clay.2016.10.021. Lee, S., van Riessen, A., Chon, C.M., Kang, N.H., Jou, H.T., Kim, Y.J., 2016. Impact of
Hu, W., Nie, Q., Huang, B., Shu, X., He, Q., 2018. Mechanical and microstructural activator type on the immobilisation of lead in fly ash-based geopolymer.
characterization of geopolymers derived from red mud and fly ashes. J. Clean. J. Hazard Mater. 305, 59e66. https://doi.org/10.1016/j.jhazmat.2015.11.023.
Prod. 186, 799e806. https://doi.org/10.1016/j.jclepro.2018.03.086. Lee, N.K., Khalid, H.R., Lee, H.K., 2017. Adsorption characteristics of cesium onto
Hu, W., Nie, Q., Huang, B., Su, A., Du, Y., Shu, X., He, Q., 2018. Mechanical property mesoporous geopolymers containing nano-crystalline zeolites. Microporous
and microstructure characteristics of geopolymer stabilized aggregate base. Mesoporous Mater. 242, 238e244. https://doi.org/10.1016/
Constr. Build. Mater. 191, 1120e1127. https://doi.org/10.1016/ j.micromeso.2017.01.030.
j.conbuildmat.2018.10.081. Li, W., Lemougna, P.N., Wang, K., He, Y., Tong, Z., Cui, X., 2017. Effect of vacuum
Huang, Y., Han, M., Yi, R., 2012. Microstructure and properties of fly ash-based dehydration on gel structure and properties of metakaolin-based geopolymers.
geopolymeric material with 5A zeolite as a filler. Constr. Build. Mater. 33, Ceram. Int. 43, 14340e14346. https://doi.org/10.1016/j.ceramint.2017.07.190.
84e89. https://doi.org/10.1016/j.conbuildmat.2012.01.014. Liew, Y.M., Kamarudin, H., Mustafa Al Bakri, A.M., Bnhussain, M., Luqman, M.,
Huang, X., Zhuang, R.L., Muhammad, F., Yu, L., Shiau, Y.C., Li, D., 2017. Solidification/ Khairul Nizar, I., Ruzaidi, C.M., Heah, C.Y., 2012. Optimization of solids-to-liquid
stabilization of chromite ore processing residue using alkali-activated com- and alkali activator ratios of calcined kaolin geopolymeric powder. Constr.
posite cementitious materials. Chemosphere 168, e221. https://doi.org/10.1016/ Build. Mater. 37, 440e451. https://doi.org/10.1016/j.conbuildmat.2012.07.075.
j.chemosphere.2016.10.067. Liu, Y., Yan, C., Qiu, X., Li, D., Wang, H., Alshameri, A., 2016. Preparation of faujasite
Huang, X., Muhammad, F., Yu, L., Jiao, B., Shiau, Y.C., Li, D., 2018. Reduction/ block from fly ash-based geopolymer via in-situ hydrothermal method.
immobilization of chromite ore processing residue using composite materials J. Taiwan Inst. Chem. Eng. 59, 433e439. https://doi.org/10.1016/
based geopolymer coupled with zero-valent iron. Ceram. Int. 44, 3454e3463. j.jtice.2015.07.012.
https://doi.org/10.1016/j.ceramint.2017.11.148. Liu, Y., Yan, C., Zhang, Z., Wang, H., Zhou, S., Zhou, W., 2016. A comparative study on
Hwang, C.L., Huynh, T.P., 2015. Effect of alkali-activator and rice husk ash content on fly ash, geopolymer and faujasite block for Pb removal from aqueous solution.
strength development of fly ash and residual rice husk ash-based geopolymers. Fuel 185, 181e189. https://doi.org/10.1016/j.fuel.2016.07.116.
Constr. Build. Mater. 101, 1e9. https://doi.org/10.1016/ Liu, Y., Yan, C., Zhang, Z., Li, L., Wang, H., Pu, S., 2017. One-step fabrication of novel
j.conbuildmat.2015.10.025. porous and permeable self-supporting zeolite block from fly ash. Mater. Lett.
Irfan Khan, M., Azizli, K., Sufian, S., Man, Z., 2015. Sodium silicate-free geopolymers 196, 328e331. https://doi.org/10.1016/j.matlet.2017.03.097.
as coating materials: effects of Na/Al and water/solid ratios on adhesion Liu, Z.H., Tang, Q., Li, C.M., He, Y., Cui, X.M., 2017. Preparation of NaA zeolite spheres
strength. Ceram. Int. 41, 2794e2805. https://doi.org/10.1016/ from geopolymer gels using a one-step method in silicone oil. Int. J. Appl.
j.ceramint.2014.10.099. Ceram. Technol. 14, 982e986. https://doi.org/10.1111/ijac.12708.
Ismail, I., Bernal, S.A., Provis, J.L., San Nicolas, R., Hamdan, S., Van Deventer, J.S.J., Lynch, J.L.V., Baykara, H., Cornejo, M., Soriano, G., Ulloa, N.A., 2018. Preparation,
2014. Modification of phase evolution in alkali-activated blast furnace slag by characterization, and determination of mechanical and thermal stability of
the incorporation of fly ash. Cement Concr. Compos. 45, 125e135. https:// natural zeolite-based foamed geopolymers. Constr. Build. Mater. 172, 448e456.
doi.org/10.1016/j.cemconcomp.2013.09.006. https://doi.org/10.1016/j.conbuildmat.2018.03.253.
Istuque, D.B., Reig, L., Moraes, J.C.B., Akasaki, J.L., Borrachero, M.V., Soriano, L., Ma, X., Zhang, Z., Wang, A., 2016. The transition of fly ash-based geopolymer gels
Paya , J., Malmonge, J.A., Tashima, M.M., 2016. Behaviour of metakaolin-based into ordered structures and the effect on the compressive strength. Constr.
geopolymers incorporating sewage sludge ash (SSA). Mater. Lett. 180, Build. Mater. 104, 25e33. https://doi.org/10.1016/j.conbuildmat.2015.12.049.
192e195. https://doi.org/10.1016/j.matlet.2016.05.137. Minelli, M., Papa, E., Medri, V., Miccio, F., Benito, P., Doghieri, F., Landi, E., 2018.
van Jaarsveld, J.G.S., van Deventer, J.S.J., Lorenzeni, L., 1997. The potential use of Characterization of novel geopolymer e zeolite composites as solid adsorbents
geopolymeric materials to immobilise toxic metals: Part 1. Theory and appli- for CO2 capture. Chem. Eng. J. 341, 505e515. https://doi.org/10.1016/
cations. Miner. Eng. 10, 659e669. j.cej.2018.02.050.
Jun, Y., Oh, J.E., 2014. Mechanical and microstructural dissimilarities in alkali- Moon, J., Bae, S., Celik, K., Yoon, S., Kim, K.H., Kim, K.S., Monteiro, P.J.M., 2014.
activation for six Class F Korean fly ashes. Constr. Build. Mater. 52, 396e403. Characterization of natural pozzolan-based geopolymeric binders. Cement
https://doi.org/10.1016/j.conbuildmat.2013.11.058. Concr. Compos. 53, 97e104. https://doi.org/10.1016/j.cemconcomp.2014.06.010.
Khalid, H.R., Lee, N.K., Park, S.M., Abbas, N., Lee, H.K., 2018. Synthesis of geopolymer- de Moraes Pinheiro, S.M., Font, A., Soriano, L., Tashima, M.M., Monzo , J.,
supported zeolites via robust one-step method and their adsorption potential. Borrachero, M.V., Paya , J., 2018. Olive-stone biomass ash (OBA): an alternative
J. Hazard Mater. 353, 522e533. https://doi.org/10.1016/j.jhazmat.2018.04.049. alkaline source for the blast furnace slag activation. Constr. Build. Mater. 178,
Khater, H.M., Abd El Gawaad, H.A., 2016. Characterization of alkali activated geo- 327e338. https://doi.org/10.1016/j.conbuildmat.2018.05.157.
polymer mortar doped with MWCNT. Constr. Build. Mater. 102, 329e337. Mozgawa, W., 2001. The relation between structure and vibrational spectra of
https://doi.org/10.1016/j.conbuildmat.2015.10.121. natural zeolites. J. Mol. Struct. 596, 129e137. https://doi.org/10.1016/S0022-
Kolousek, D., Brus, J., Urbanova, M., Andertova, J., Hulinsky, V., Vorel, J., 2007. 2860(01)00741-4.
Preparation, structure and hydrothermal stability of alternative (sodium Muhammad, F., Huang, X., Li, S., Xia, M., Zhang, M., Liu, Q., Shehzad Hassan, M.A.,
silicate-free) geopolymers. J. Mater. Sci. 42, 9267e9275. https://doi.org/10.1007/ Jiao, B., Yu, L., Li, D., 2018. Strength evaluation by using polycarboxylate
s10853-007-1910-5. superplasticizer and solidification efficiency of Cr6þ, Pb2þand Cd2þin com-
Komljenovi c, M., Bas
carevi
c, Z., Bradi
c, V., 2010. Mechanical and microstructural posite based geopolymer. J. Clean. Prod. 188, 807e815. https://doi.org/10.1016/
properties of alkali-activated fly ash geopolymers. J. Hazard Mater. 181, 35e42. j.jclepro.2018.04.033.
https://doi.org/10.1016/j.jhazmat.2010.04.064. Naghsh, M., Shams, K., 2017. Synthesis of a kaolin-based geopolymer using a novel
Komnitsas, K., Zaharaki, D., Perdikatsis, V., 2009. Effect of synthesis parameters on fusion method and its application in effective water softening. Appl. Clay Sci.
the compressive strength of low-calcium ferronickel slag inorganic polymers. 146, 238e245. https://doi.org/10.1016/j.clay.2017.06.008.
J. Hazard Mater. 161, 760e768. https://doi.org/10.1016/j.jhazmat.2008.04.055. Najafi Kani, E., Allahverdi, A., Provis, J.L., 2012. Efflorescence control in geopolymer
578 P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579

binders based on natural pozzolan. Cement Concr. Compos. 34, 25e33. https:// Prud'Homme, E., Michaud, P., Joussein, E., Rossignol, S., 2012. Influence of raw
doi.org/10.1007/s10973-016-5850-7. materials and potassium and silicon concentrations on the formation of a
Nath, S.K., Maitra, S., Mukherjee, S., Kumar, S., 2016. Microstructural and morpho- zeolite phase in a geopolymer network during thermal treatment. J. Non-Cryst.
logical evolution of fly ash based geopolymers. Constr. Build. Mater. 111, Solids 358, 1908e1916. https://doi.org/10.1016/j.jnoncrysol.2012.05.043.
758e765. https://doi.org/10.1016/j.conbuildmat.2016.02.106. Qiu, X., Liu, Y., Li, D., Yan, C., 2014. Preparation of NaP zeolite block from fly ash-
Nazari, A., Bagheri, A., Dao, M., Mallawa, C., Zannis, P., Zumbo, S., Sanjayan, J.G., 2017. based geopolymer via in situ hydrothermal method. J. Porous Mater. 22,
The behaviour of iron in geopolymer under thermal shock. Constr. Build. Mater. 291e299. https://doi.org/10.1007/s10934-014-9895-3.
150, 248e251. https://doi.org/10.1016/j.conbuildmat.2017.05.223. Rasouli, H.R., Golestani-Fard, F., Mirhabibi, A.R., Nasab, G.M., MacKenzie, K.J.D.,
Nguyen, A.D., Skv ara, F., 2016. The influence of ambient pH on fly ash-based geo- Shahraki, M.H., 2015. Fabrication and properties of microporous metakaolin-
polymer. Cement Concr. Compos. 72, 275e283. https://doi.org/10.1016/ based geopolymer bodies with polylactic acid (PLA) fibers as pore generators.
j.cemconcomp.2016.06.010. Ceram. Int. 41, 7872e7880. https://doi.org/10.1016/j.ceramint.2015.02.125.
Nie, Q., Hu, W., Huang, B., Shu, X., He, Q., 2019. Synergistic utilization of red mud for Rees, C.A., Provis, J.L., Lukey, G.C., van Deventer, J.S., 2007. Attenuated total reflec-
flue-gas desulfurization and fly ash-based geopolymer preparation. J. Hazard tance Fourier transform infrared analysis of fly ash geopolymer gel aging.
Mater. 369, 503e511. https://doi.org/10.1016/j.jhazmat.2019.02.059. Langmuir 23, 8170e8179. https://doi.org/10.1021/la700713g.
Nikolov, A., Rostovsky, I., Nugteren, H., 2017. Geopolymer materials based on natural Rees, C.A., Provis, J.L., Lukey, G.C., van Deventer, J.S.J., 2008. The mechanism of
zeolite. Case Stud. Constr. Mater. 6, 198e205. https://doi.org/10.1016/ geopolymer gel formation investigated through seeded nucleation. Colloid.
j.cscm.2017.03.001. Surf. Physicochem. Eng. Asp. 318, 97e105. https://doi.org/10.1016/
Noor-Ul-Amin, Faisal, M., Muhammad, K., Gul, S., 2016. Synthesis and character- j.colsurfa.2007.12.019.
ization of geopolymer from bagasse bottom ash, waste of sugar industries and Rodríguez, E.D., Bernal, S.A., Provis, J.L., Gehman, J.D., Monzo , J.M., Paya, J.,
naturally available China clay. J. Clean. Prod. 129, 491e495. https://doi.org/ Borrachero, M.V., 2013. Geopolymers based on spent catalyst residue from a
10.1016/j.jclepro.2016.04.024. fluid catalytic cracking (FCC) process. Fuel 109, 493e502. https://doi.org/
Ofer-Rozovsky, E., Arbel Haddad, M., Bar Nes, G., Katz, A., 2016. The formation of 10.1016/j.fuel.2013.02.053.
crystalline phases in metakaolin-based geopolymers in the presence of sodium Rodríguez, E.D., Bernal, S.A., Provis, J.L., Paya, J., Monzo, J.M., Borrachero, M.V., 2013.
nitrate. J. Mater. Sci. 51, 4795e4814. https://doi.org/10.1007/s10853-016-9767- Effect of nanosilica-based activators on the performance of an alkali-activated
0. fly ash binder. Cement Concr. Compos. 35, 1e11. https://doi.org/10.1016/
Oh, J.E., Monteiro, P.J.M., Jun, S.S., Choi, S., Clark, S.M., 2010. The evolution of j.cemconcomp.2012.08.025.
strength and crystalline phases for alkali-activated ground blast furnace slag De Rossi, A., Sim~ ao, L., Ribeiro, M.J., Novais, R.M., Labrincha, J.A., Hotza, D.,
and fly ash-based geopolymers. Cement Concr. Res. 40, 189e196. https:// Moreira, R.F.P.M., 2018. In-situ synthesis of zeolites by geopolymerization of
doi.org/10.1016/j.cemconres.2009.10.010. biomass fly ash and metakaolin. Mater. Lett. https://doi.org/10.1016/
Oh, J.E., Moon, J., Mancio, M., Clark, S.M., Monteiro, P.J.M., 2011. Bulk modulus of j.matlet.2018.11.016.
basic sodalite, Na8[AlSiO4] 6(OH)2 2H2O, a possible zeolitic precursor in coal- Roy, D.M., 1999. Alkali activated cements, opportunities and challenges. Cement
fly-ash-based geopolymers. Cement Concr. Res. 41, 107e112. https://doi.org/ Concr. Res. 29, 249e254.
10.1016/j.cemconres.2010.09.012. _
Rozek, l, M., Mozgawa, W., 2018. Spectroscopic studies of fly ash-based
P., Kro
Oh, J.E., Clark, S.M., Monteiro, P.J.M., 2011. Determination of the bulk modulus of geopolymers. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 198, 283e289.
hydroxycancrinite, a possible zeolitic precursor in geopolymers, by high- https://doi.org/10.1016/j.saa.2018.03.034.
pressure synchrotron X-ray diffraction. Cement Concr. Compos. 33, Shearer, C.R., Provis, J.L., Bernal, S.A., Kurtis, K.E., 2016. Alkali-activation potential of
1014e1019. https://doi.org/10.1016/j.cemconcomp.2011.05.002. biomass-coal co-fired fly ash. Cement Concr. Compos. 73, 62e74. https://
Oh, J.E., Jun, Y., Jeong, Y., 2014. Characterization of geopolymers from composi- doi.org/10.1016/j.cemconcomp.2016.06.014.
tionally and physically different Class F fly ashes. Cement Concr. Compos. 50, Shi, Z., Shi, C., Zhang, J., Wan, S., Zhang, Z., Ou, Z., 2018. Alkali-silica reaction in
16e26. https://doi.org/10.1016/j.cemconcomp.2013.10.019. waterglass-activated slag mortars incorporating fly ash and metakaolin.
Ozer, I., Soyer-Uzun, S., 2015. Relations between the structural characteristics and Cement Concr. Res. 108, 10e19. https://doi.org/10.1016/
compressive strength in metakaolin based geopolymers with different molar Si/ j.cemconres.2018.03.002.
Al ratios. Ceram. Int. 41, 10192e10198. https://doi.org/10.1016/ Shiota, K., Nakamura, T., Takaoka, M., Aminuddin, S.F., Oshita, K., Fujimori, T., 2017.
j.ceramint.2015.04.125. Stabilization of cesium in alkali-activated municipal solid waste incineration fly
Pacheco-Torgal, F., Castro-Gomes, J., Jalali, S., 2008. Alkali-activated binders: a re- ash and a pyrophyllite-based system. Chemosphere 187, 188e195. https://
view: Part 1. Historical background, terminology, reaction mechanisms and doi.org/10.1016/j.chemosphere.2017.08.114.
hydration products. Constr. Build. Mater. 22, 1305e1314. https://doi.org/ Shiota, K., Nakamura, T., Takaoka, M., Aminuddin, S.F., Oshita, K., Fujimori, T., 2017.
10.1016/j.conbuildmat.2007.10.015. Stabilization of lead in an alkali-activated municipal solid waste incineration fly
Palomo, A., Alonso, S., Ferna ndez-Jime nez, A., 2008. Alkaline Activation of Fly ashePyrophyllite-based system. J. Environ. Manag. 201, 327e334. https://
Ashes : NMR Study of the Reaction Products Alkaline Activation of Fly Ashes : doi.org/10.1016/j.jenvman.2017.07.002.
NMR Study of the Reaction Products, vol. 1145, pp. 1141e1145. https://doi.org/ De Silva, P., Sagoe-Crenstil, K., 2008. Medium-term phase stability of Na2O-Al2O3-
10.1111/j.1551-2916.2004.01141.x. J. SiO2-H2O geopolymer systems. Cement Concr. Res. 38, 870e876. https://
Panagiotopoulou, C., Tsivilis, S., Kakali, G., 2015. Application of the Taguchi approach doi.org/10.1016/j.cemconres.2007.10.003.
for the composition optimization of alkali activated fly ash binders. Constr. Silva, J.S., Medeiros De Jesus-Neto, R., Fiuza, R.A., Gonçalves, J.P., Mascarenhas, A.J.S.,
Build. Mater. 91, 17e22. https://doi.org/10.1016/j.conbuildmat.2015.05.005. Andrade, H.M.C., 2016. Alkali-activation of spent fluid cracking catalysts for CO2
Papa, E., Medri, V., Amari, S., Benito, P., Vaccari, A., Landi, E., 2018. Zeolite-geo- capture. Microporous Mesoporous Mater. 232, 1e12. https://doi.org/10.1016/
polymer composite materials : production and characterization. J. Clean. Prod. j.micromeso.2016.06.005.
171, 76e84. https://doi.org/10.1016/j.jclepro.2017.09.270. Siyal, A.A., Shamsuddin, M.R., Khan, M.I., Rabat, N.E., Zulfiqar, M., Man, Z., Siame, J.,
Peng, M.X., Wang, Z.H., Shen, S.H., Xiao, Q.G., Li, L.J., Tang, Y.C., Hu, L.L., 2017. Alkali Azizli, K.A., 2018. A review on geopolymers as emerging materials for the
fusion of bentonite to synthesize one-part geopolymeric cements cured at adsorption of heavy metals and dyes. J. Environ. Manag. 224, 327e339. https://
elevated temperature by comparison with two-part ones. Constr. Build. Mater. doi.org/10.1016/j.jenvman.2018.07.046.
130, 103e112. https://doi.org/10.1016/j.conbuildmat.2016.11.010. Slaty, F., Khoury, H., Wastiels, J., Rahier, H., 2013. Characterization of alkali activated
Peng, M.X., Wang, Z.H., Xiao, Q.G., Song, F., Xie, W., Yu, L.C., Huang, H.W., Yi, S.J., kaolinitic clay. Appl. Clay Sci. 75e76, 120e125. https://doi.org/10.1016/
2017. Effects of alkali on one-part alkali-activated cement synthesized by j.clay.2013.02.005.
calcining bentonite with dolomite and Na2CO3. Appl. Clay Sci. 139, 64e71. Sore, S.O., Messan, A., Prud’homme, E., Escadeillas, G., Tsobnang, F., 2016. Synthesis
https://doi.org/10.1016/j.clay.2017.01.020. and characterization of geopolymer binders based on local materials from
Provis, J.L., van Deventer, J.S.J., 2007. Geopolymerisation kinetics. 1. In situ energy- Burkina Faso e metakaolin and rice husk ash. Constr. Build. Mater. 124,
dispersive X-ray diffractometry. Chem. Eng. Sci. 62, 2309e2317. https:// 301e311. https://doi.org/10.1016/j.conbuildmat.2016.07.102.
doi.org/10.1016/j.ces.2007.01.027. Sturm, P., Gluth, G.J.G., Simon, S., Brouwers, H.J.H., Kühne, H.C., 2016. The effect of
Provis, J.L., van Deventer, J.S.J., 2007. Geopolymerisation kinetics. 2. Reaction kinetic heat treatment on the mechanical and structural properties of one-part geo-
modelling. Chem. Eng. Sci. 62, 2318e2329. https://doi.org/10.1016/ polymer-zeolite composites. Thermochim. Acta 635, 41e58. https://doi.org/
j.ces.2007.01.028. 10.1016/j.tca.2016.04.015.
Provis, J.L., Van Deventer, J.S.J. (Eds.), 2009. Geopolymers: Structures, Processing, Sturm, P., Gluth, G.J.G., J€ ager, C., Brouwers, H.J.H., Kühne, H.C., 2018. Sulfuric acid
Properties and Industrial Applications. Elsevier. resistance of one-part alkali-activated mortars. Cement Concr. Res. 109, 54e63.
Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2005. Do geopolymers actually contain https://doi.org/10.1016/j.cemconres.2018.04.009.
nanocrystalline zeolites? A reexamination of existing results. Chem. Mater. 17, Sudagar, A., Andrejkovi , S., Patinha, C., Velosa, A., McAdam, A., da Silva, E.F.,
cova
3075e3085. https://doi.org/10.1021/cm050230i. Rocha, F., 2018. A novel study on the influence of cork waste residue on
Provis, J.L., Yong, C.Z., Duxson, P., van Deventer, J.S.J., 2009. Correlating mechanical metakaolin-zeolite based geopolymers. Appl. Clay Sci. 152, 196e210. https://
and thermal properties of sodium silicate-fly ash geopolymers. Colloid. Surf. doi.org/10.1016/j.clay.2017.11.013.
Physicochem. Eng. Asp. 336, 57e63. https://doi.org/10.1016/ Sukmak, P., Horpibulsuk, S., Shen, S.L., 2013. Strength development in clay-fly ash
j.colsurfa.2008.11.019. geopolymer. Constr. Build. Mater. 40, 566e574. https://doi.org/10.1016/
Prud'Homme, E., Michaud, P., Joussein, E., Clacens, J.M., Arii-Clacens, S., Sobrados, I., j.conbuildmat.2012.11.015.
Peyratout, C., Smith, A., Sanz, J., Rossignol, S., 2011. Structural characterization of Suksiripattanapong, C., Horpibulsuk, S., Chanprasert, P., Sukmak, P., Arulrajah, A.,
geomaterial foams - thermal behavior. J. Non-Cryst. Solids 357, 3637e3647. 2015. Compressive strength development in fly ash geopolymer masonry units
https://doi.org/10.1016/j.jnoncrysol.2011.06.033. manufactured from water treatment sludge. Constr. Build. Mater. 82, 20e30.
P. Roz_ ek et al. / Journal of Cleaner Production 230 (2019) 557e579 579

https://doi.org/10.1016/j.conbuildmat.2015.02.040. Xu, H., Gong, W., Syltebo, L., Izzo, K., Lutze, W., Pegg, I.L., 2014. Effect of blast furnace
Sun, Z., Vollpracht, A., 2018. Isothermal calorimetry and in-situ XRD study of the slag grades on fly ash based geopolymer waste forms. Fuel 133, 332e340.
NaOH activated fly ash, metakaolin and slag. Cement Concr. Res. 103, 110e122. https://doi.org/10.1016/j.fuel.2014.05.018.
https://doi.org/10.1016/j.cemconres.2017.10.004. Xu, Z., Jiang, Z., Wu, D., Peng, X., Xu, Y., Li, N., Qi, Y., Li, P., 2016. Immobilization of
Takeda, H., Hashimoto, S., Yokoyama, H., Honda, S., Iwamoto, Y., 2013. Character- strontium-loaded zeolite A by metakaolin based-geopolymer. Ceram. Int. 43,
ization of zeolite in zeolite-geopolymer hybrid bulk materials derived from 4434e4439. https://doi.org/10.1016/j.ceramint.2016.12.092.
kaolinitic clays. Materials 6, 1767e1778. https://doi.org/10.3390/ma6051767. Xu, M., He, Y., Wang, Y., Cui, X., 2017. Preparation of a non-hydrothermal NaA zeolite
Tang, Q., Ge, Y.Y., Wang, K.T., He, Y., Cui, X.M., 2015. Preparation of porous P-type membrane and defect elimination by vacuum-inhalation repair method. Chem.
zeolite spheres with suspension solidification method. Mater. Lett. 161, Eng. Sci. 158, 117e123. https://doi.org/10.1016/j.ces.2016.10.001.
558e560. https://doi.org/10.1016/j.matlet.2015.09.062. Xuemin, C., Yan, H., Leping, L., Jinyu, C., 2011. NaA zeolite synthesis from geopolymer
Tang, Q., He, Y., pin Wang, Y., tuo Wang, K., min Cui, X., 2016. Study on synthesis and precursor. MRS Commun. 1, 49e51. https://doi.org/10.1557/mrc.2011.15.
characterization of ZSM-20 zeolites from metakaolin-based geopolymers. Appl. Yan, H., Xue-Min, C., Jin, M., Liu, L.P., Liu, X.D., Chen, J.Y., 2012. The hydrothermal
Clay Sci. 129, 102e107. https://doi.org/10.1016/j.clay.2016.05.011. transformation of solid geopolymers into zeolites. Microporous Mesoporous
Temuujin, J., Minjigmaa, A., Rickard, W., Lee, M., Williams, I., van Riessen, A., 2009. Mater. 161, 187e192. https://doi.org/10.1016/j.micromeso.2012.05.039.
Preparation of metakaolin based geopolymer coatings on metal substrates as Ye, H., Radlin ska, A., 2016. Fly ash-slag interaction during alkaline activation: in-
thermal barriers. Appl. Clay Sci. 46, 265e270. https://doi.org/10.1016/ fluence of activators on phase assemblage and microstructure formation.
j.clay.2009.08.015. Constr. Build. Mater. 122, 594e606. https://doi.org/10.1016/
Temuujin, J., Minjigmaa, A., Rickard, W., Lee, M., Williams, I., van Riessen, A., 2010. j.conbuildmat.2016.06.099.
Fly ash based geopolymer thin coatings on metal substrates and its thermal Ye, N., Yang, J., Ke, X., Zhu, J., Li, Y., Xiang, C., Wang, H., Li, L., Xiao, B., 2014. Synthesis
evaluation. J. Hazard Mater. 180, 748e752. https://doi.org/10.1016/ and characterization of geopolymer from Bayer red mud with thermal pre-
j.jhazmat.2010.04.121. treatment. J. Am. Ceram. Soc. 97, 1652e1660. https://doi.org/10.1111/jace.12840.
Temuujin, J., Minjigmaa, A., Davaabal, B., Bayarzul, U., Ankhtuya, A., Jadambaa, T., Ye, N., Yang, J., Liang, S., Hu, Y., Hu, J., Xiao, B., Huang, Q., 2016. Synthesis and
Mackenzie, K.J.D., 2014. Utilization of radioactive high-calcium Mongolian fly strength optimization of one-part geopolymer based on red mud. Constr. Build.
ash for the preparation of alkali-activated geopolymers for safe use as con- Mater. 111, 317e325. https://doi.org/10.1016/j.conbuildmat.2016.02.099.
struction materials. Ceram. Int. 40, 16475e16483. https://doi.org/10.1016/ Yousef, R.I., El-Eswed, B., Alshaaer, M., Khalili, F., Khoury, H., 2009. The influence of
j.ceramint.2014.07.157. using Jordanian natural zeolite on the adsorption, physical, and mechanical
Torres-Carrasco, M., Puertas, F., 2015. Waste glass in the geopolymer preparation. properties of geopolymers products. J. Hazard Mater. 165, 379e387. https://
Mechanical and microstructural characterisation. J. Clean. Prod. 90, 397e408. doi.org/10.1016/j.jhazmat.2008.10.004.
https://doi.org/10.1016/j.jclepro.2014.11.074. Yousef, R.I., El-Eswed, B., Alshaaer, M., Khalili, F., Rahier, H., 2012. Degree of reac-
Tyni, S.K., Karppinen, J.A., Tiainen, M.S., Laitinen, R.S., 2014. Preparation and char- tivity of two kaolinitic minerals in alkali solution using zeolitic tuff or silica
acterization of amorphous aluminosilicate polymers from ash formed in com- sand filler. Ceram. Int. 38, 5061e5067. https://doi.org/10.1016/
bustion of peat and wood mixtures. J. Non-Cryst. Solids 387, 94e100. https:// j.ceramint.2012.03.008.
doi.org/10.1016/j.jnoncrysol.2013.12.032. Yun-Ming, L., Cheng-Yong, H., Li, L., Jaya, N.A., Abdullah, M.M.A.B., Jin, T.S.,
Tzanakos, K., Mimilidou, A., Anastasiadou, K., Stratakis, A., Gidarakos, E., 2014. So- Hussin, K., 2017. Formation of one-part-mixing geopolymers and geopolymer
lidification/stabilization of ash from medical waste incineration into geo- ceramics from geopolymer powder. Constr. Build. Mater. 156, 9e18. https://
polymers. Waste Manag. 34, 1823e1828. https://doi.org/10.1016/ doi.org/10.1016/j.conbuildmat.2017.08.110.
j.wasman.2014.03.021. Zhang, J., Provis, J.L., Feng, D., van Deventer, J.S.J., 2008. Geopolymers for immobi-
Ulloa, N.A., Baykara, H., Cornejo, M.H., Rigail, A., Paredes, C., Luis, J., 2018. Appli- lization of Cr6þ, Cd2þ, and Pb2þ. J. Hazard Mater. 157, 587e598. https://
cation-oriented mix design optimization and characterization of zeolite-based doi.org/10.1016/j.jhazmat.2008.01.053.
geopolymer mortars. Constr. Build. Mater. 174, 138e149. https://doi.org/ Zhang, Y.J., Wang, Y.C., Xu, D.L., Li, S., 2010. Mechanical performance and hydration
10.1016/j.conbuildmat.2018.04.101. mechanism of geopolymer composite reinforced by resin. Mater. Sci. Eng. A 527,
Vaiciukyniene, _ D., Borg, R.P., Kiele,
_ A., Kantautas, A., 2018. Alkali-activated blends of 6574e6580. https://doi.org/10.1016/j.msea.2010.06.069.
calcined AlF3 production waste and clay. Ceram. Int. 44, 12573e12579. https:// Zhang, Z., Wang, H., Provis, J.L., Bullen, F., Reid, A., Zhu, Y., 2012. Quantitative kinetic
doi.org/10.1016/j.ceramint.2018.04.054. and structural analysis of geopolymers. Part 1. the activation of metakaolin with
mez, L., 2010. Geopolymer synthesis using alkaline
Villa, C., Pecina, E.T., Torres, R., Go sodium hydroxide. Thermochim. Acta 539, 23e33. https://doi.org/10.1016/
activation of natural zeolite. Constr. Build. Mater. 24, 2084e2090. https:// j.tca.2012.03.021.
doi.org/10.1016/j.conbuildmat.2010.04.052. Zhang, J., He, Y., Wang, Y.P., Mao, J., Cui, X.M., 2014. Synthesis of a self-Supporting
Waijarean, N., MacKenzie, K.J.D., Asavapisit, S., Piyaphanuwat, R., Jameson, G.N.L., faujasite zeolite membrane using geopolymer gel for separation of alcohol/
2017. Synthesis and properties of geopolymers based on water treatment res- water mixture. Mater. Lett. 116, 167e170. https://doi.org/10.1016/
idue and their immobilization of some heavy metals. J. Mater. Sci. 52, j.matlet.2013.11.008.
7345e7359. https://doi.org/10.1007/s10853-017-0970-4. Zhang, Z., Provis, J.L., Reid, A., Wang, H., 2015. Mechanical, thermal insulation,
Wan, Q., Rao, F., Song, S., García, R.E., Estrella, R.M., Patin ~ o, C.L., Zhang, Y., 2017. thermal resistance and acoustic absorption properties of geopolymer foam
Geopolymerization reaction, microstructure and simulation of metakaolin- concrete. Cement Concr. Compos. 62, 97e105. https://doi.org/10.1016/
based geopolymers at extended Si/Al ratios. Cement Concr. Compos. 79, j.cemconcomp.2015.03.013.
45e52. https://doi.org/10.1016/j.cemconcomp.2017.01.014. Zhang, Z., Li, L., He, D., Ma, X., Yan, C., Wang, H., 2016. Novel self-supporting zeolitic
Wan, Q., Rao, F., Song, S., García, R.E., Estrella, R.M., Patin ~ o, C.L., Zhang, Y., 2017. block with tunable porosity and crystallinity for water treatment. Mater. Lett.
Geopolymerization reaction, microstructure and simulation of metakaolin- 178, 151e154. https://doi.org/10.1016/j.matlet.2016.04.214.
based geopolymers at extended Si/Al ratios. Cement Concr. Compos. 79, Zhang, P., Zheng, Y., Wang, K., Zhang, J., 2018. A review on properties of fresh and
45e52. https://doi.org/10.1016/j.cemconcomp.2017.01.014. hardened geopolymer mortar. Compos. B Eng. 152, 79e95. https://doi.org/
Wang, J., Wang, J.X., Zhang, Q., Li, Y.X., 2016. Immobilization of simulated low and 10.1016/j.compositesb.2018.06.031.
intermediate level waste in alkali-activated slag-fly ash-metakaolin hydro- Zhang, Y.J., Chen, H., He, P.Y., Juan Li, C., 2019. Developing silica fume-based self-
ceramics. Nucl. Eng. Des. 300, 67e73. https://doi.org/10.1016/ supported ECR-1 zeolite membrane for seawater desalination. Mater. Lett. 236,
j.nucengdes.2016.01.011. 538e541. https://doi.org/10.1016/j.matlet.2018.10.167.
Wang, H., Yan, C., Li, D., Zhou, F., Liu, Y., Zhou, C., 2018. In situ transformation of Zhu, H., Zhang, Z., Deng, F., Cao, Y., 2013. The effects of phase changes on the
geopolymer gels to self-supporting NaX zeolite monoliths with excellent bonding property of geopolymer to hydrated cement. Constr. Build. Mater. 48,
compressive strength. Microporous Mesoporous Mater. 261, 164e169. https:// 124e130. https://doi.org/10.1016/j.conbuildmat.2013.06.095.
doi.org/10.1016/j.micromeso.2017.11.015. Zibouche, F., Kerdjoudj, H., de Lacaillerie, J.B. d. E., Van Damme, H., 2009. Geo-
Xu, H., Van Deventer, J.S.J., 2003. The effect of alkali metals on the formation of polymers from Algerian metakaolin. Influence of secondary minerals. Appl. Clay
geopolymeric gels from alkali-feldspars. Colloid. Surf. Physicochem. Eng. Asp. Sci. 43, 453e458. https://doi.org/10.1016/j.clay.2008.11.001.
216, 27e44. https://doi.org/10.1016/S0927-7757(02)00499-5.

You might also like