Valorization of an Electronic Waste-Derived Aluminosilicate: Surface

May 3, 2016 - Valorization of an Electronic Waste-Derived Aluminosilicate: Surface Functionalization and Porous Structure Tuning. Chao Ning†, Pejman...
1 downloads 6 Views 2MB Size
Subscriber access provided by Brown University Library

Article

Valorization of an electronic waste-derived aluminosilicate – Surface functionalization and porous structure tuning Chao Ning, Pejman Hadi, Carol Sze Ki Lin, and Gordon McKay ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.5b01523 • Publication Date (Web): 03 May 2016 Downloaded from http://pubs.acs.org on May 7, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Valorization of an electronic waste-derived aluminosilicate – Surface functionalization and porous structure tuning

Chao Ninga, Pejman Hadia, Carol Sze Ki Linb and Gordon McKaya,c*

a

Chemical and Biomolecular Engineering Department, Hong Kong University of Science and

Technology, Clear Water Bay Road, Hong Kong SAR b

School of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Hong

Kong SAR c

Division of Sustainable Development, College of Science, Engineering and Technology, Hamad

Bin Khalifa University, Qatar Foundation, Doha, Qatar *

Tel: +852 23588412, Fax: +852 23580054, E-mail: [email protected]

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract This study involves the sustainable development of an ion exchange material with ultra-high heavy metal uptake capacity from a waste material, originally destined for landfills. In this study, a promising thermo-alkaline reaction has been employed to simultaneously alter the surface chemistry and tune the textural properties of the waste-derived aluminosilicate. The effects of several reaction variables on the formation of mesotunnels in the structure of the material have been examined. Also, the surface characterization of the functionalized aluminosilicate has demonstrated that the functionalization reaction results in the cleavage of the robust  −  −   linkages (where     = ) into  −  moieties, counter-balanced by an alkali metal cation, resulting in the coverage of the aluminosilicate surface with active ion exchange sites. Comparison of the ion exchange capacity of the functionalized aluminosilicate with the commercial ion exchange resins has proved exceptionally higher heavy metal uptake for the former. The ultra-high heavy metal uptake of this material is ascribed to the high concentration of developed counter-balancing cations on the material surface. The attractiveness of this innovative approach is manifested by the dual environmental benefit, i.e. sustainable upcycling of a waste formerly disposed of into landfills and its utilization for heavy metal-laden wastewater treatment.

Keywords: Sustainable development; Waste – derived aluminosilicate; Mesoporous structure; Ion exchange; Heavy metal removal; Adsorption; Functionalization

ACS Paragon Plus Environment

Page 2 of 46

Page 3 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Introduction Aluminosilicate materials have received ever-increasing attention due to their high internal surface area and pore volume, tunable pore size distribution, high thermal stability and unique ion exchange capability1. Fabrication of these materials with desired textural properties and their surface modification to tailor certain functionalities stand out as an utmost crucial area of research2, as they find a wide array of applications in electrochemistry3, catalysis4–7, support8,9, cement industry10, chromatography11, dehumidification12, antioxidant13, fertilizers14, sensors15 and sorbent16–19. Ion exchange has been one of the main areas of interest for aluminosilicate materials for catalysis20 and adsorption21. Several investigations have been carried out for the isomorphous substitution of trivalent aluminum by tetravalent silicon in the aluminosilicate skeleton during the synthesis process giving rise to the net negative charge on its surface. This negativelycharged framework is balanced with extra-framework alkali metal cations22,23. These cations can theoretically be exchanged with target metal species, such as heavy metals, either to tailor certain other properties to the aluminosilicate (such as in catalysis)24 or to remove charged pollutants (including heavy metal ions) from a medium (such as in wastewater treatment)25. With the recent advent of environmental awareness, numerous researchers have attempted to sustainably exploit the ion exchange ability of natural aluminosilicate-based minerals, such as clinoptilolite26,27, kaolin28,29 and mordenite30, as an alternative for unsustainable multi-stage production of synthetic aluminosilicate requiring the reckless utilization of material and energy resources. Although some progress has been made in this area, the inherent drawback with this so-called sustainable concept is the low ion exchange capacity of these minerals31, which hinders their commercial utilization in applications requiring high ion exchange capacity. Gupta and

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bhattacharyya have summarized the ion exchange capacities of many different types of inorganic materials32. Their tabulated results show the extremely low metal uptake by natural inorganic minerals. Therefore, it is deemed necessary to develop new sustainable approaches to fabricate cost-effective ion exchange materials without sacrificing the efficiency. As part of our ongoing investigations, we believe that the low ion exchange capacity of aluminosilicates originates from the fact that the alkali cations, functioning as a chargecompensator for   clusters, can only locally move along the lattice and hence their potential to actively participate in ion exchange reactions is masked by two phenomena: barrier against the diffusion of the alkali metal cations through the pores to be exchanged with heavy metals 1, and resistance against the electroneutrality disturbance in the course of ion exchange. Therefore, we ascribe the low ion exchange capacity of the natural and synthetic zeolites predominantly to the microporous nature of their architecture as well as the great electrostatic interaction between the charge-balancing metal cations and the zeolite structure. Alternatively, the alkali cations in the position of network modifier, bound to the non-bridging oxygens (), are more likely to play the role of ion exchange sites because of the small diffusion barrier and comparatively low electrostatic attraction33. Yet, in most of the studies, this distinction between the charge compensator cations and network modifier species has been disregarded. In the light of this hypothesis, increasing the number of  and thus ion exchange sites is a more promising approach to boost the ion exchange capability of an aluminosilicate-based material. The aim of this study is to create mesotunnels covered with active ion exchange sites in a hydrophobic nonporous aluminosilicate material derived from an electronic waste source and transform it into a functionalized hydrophilic ion exchange material. Recently, there has been an enormous outcry on the environmental impacts of waste printed

ACS Paragon Plus Environment

Page 4 of 46

Page 5 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

circuit boards (PCBs) 34,35. Since waste PCBs comprise a myriad of heavy metals, they cannot be disposed of into landfills 36–38. In addition to the toxicity issue, the low calorific value of these materials does not allow their incineration. On the other hand, PCBs can be acknowledged as a mine for concentrated metals 39,40 as well as a resource for aluminosilicates35,41, the recycling of which can contribute to the environmental mitigation and add value to a waste substance. Hence, the sustainable development of ion exchange material from a problematic waste source will not only divert the waste disposal practice into landfills, but can also be commercially used for heavy metal-laden wastewater treatment, doubling the stimulus for greener environment. In this study, the effects of the reaction parameters on the mesotunnel formation with functionalized walls via the partial cleavage of unreactive  −  −   linkages (where  is bridging oxygen () and ,   = ) into more reactive  −  …  linkages (where  is non-bridging oxygen () and  is an alkali metal cation) using a thermo-alkaline reaction have been elucidates.

Experimental Surface modification of the NMF aluminosilicate Waste printed circuit boards (WPCBs) had been crushed into fine powders in a local recycling company and the metal and nonmetal fractions were separated by a corona electrostatic discharge process. Non-metallic fraction of WPCB (NMF), a hydrophobic powder-form material with an average particle size of 2 , was used as an inexpensive amorphous aluminosilicate source. NMF was impregnated with 1M potassium hydroxide solution (Sigma-Aldrich, >85%), as the activating agent, at a range of impregnation ratios (1 − 4 /) for 1 h until homogeneous dispersion was reached. In this context, impregnation ratio is defined as the weight ratio of the

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

chemical reagent to the precursor. After the completion of the impregnation process, the surface functionalization reaction was carried out by heating the mixture in a reactor to a desired activation temperature (200-400˚C) and for specified reaction durations (1-7 h) under an inert nitrogen atmosphere. Subsequently, the surface functionalized material was washed several times with hot water to remove excessive unreacted activating agent, freeze dried overnight and kept in a desiccator for later use. Notably, minimum of 3 replications for each reaction were conducted to ascertain the accuracy of the results.

Surface characterization The surface morphologies of the virgin and functionalized aluminosilicate were investigated by a JEOL JSM-6390 scanning electron microscopy (SEM) coupled with energy-dispersive X-ray spectroscopy (EDX). The samples were mounted on a carbon tape and coated with a thin layer of gold via sputtering technique. Random areas of the sample surfaces were selected for EDX elementary mapping. Fourier transform infrared spectroscopy (FTIR) was recorded on a FTS 6000 FTIR spectrometer with a nominal resolution of 2  at a range of 4000-400  for both virgin and functionalized samples to determine their surface functional groups. X-ray photoelectron spectroscopy (XPS, model PHI5600) was conducted with a monochromatic Al Kα source (excitation energy, hυ = 1486.6 eV) at a voltage of 10 kV and a current of 15 mA. Low resolution spectrum for the whole range was acquired at 70 eV pass energy, whereas high resolution narrow range spectra for elements of interest were acquired at 20 eV pass energy. The water contact angle measurements were determined using GBX Digidrop (model No SERIE 2850902D).

ACS Paragon Plus Environment

Page 6 of 46

Page 7 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Textural properties Nitrogen adsorption-desorption isotherms were obtained using a Beckman Coulter SA3100 surface area analyzer at a relative pressure range of "

!

= 0.01 − 0.99 and at the liquid

nitrogen temperature (77 K). The functionalized materials were outgassed at 150˚C for 90 min in a high vacuum prior to analysis in order to remove moisture and any other impurities potentially physisorbed to the pores of the materials. Specific surface area ( &'( ) was determined at the linear relative pressure range of 0.05 – 0.3 using the Brunauer – Emmett – Teller ()) equation42. The total pore volume (*+ ) was calculated at a relative pressure of 0.98. The t-plot method43 was employed to determine the micropore volume (*-./ ). The pore size distribution was obtained using the Barrett – Joyner –Halenda (01) equation44 at the desorption branch of the isotherm. Average pore diameter (2 ) was estimated considering the cylindrical pore geometry.

Adsorption experiments The batch adsorption tests were performed by adding 50 mg ion exchange material into bottles containing 50  metal solution with a range of concentrations (0.5 − 7  . 5 ). The solution pH level was adjusted to 4 using 0.5 M HNO3 and NaOH solutions to eliminate any potential effect of acidity/basicity. The solutions were agitated in an orbital shaker at a rate of 120 rpm and a constant temperature of 25˚C until equilibrium. The initial and final metal concentrations were determined by an inductively coupled plasma atomic emission spectrometer (ICP-AES, Optima 7300DV, Perkin Elmer). The metal uptake capacities of the functionalized materials were determined using the mass balance between the liquid and solid phases.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Result and Discussion Tunneling effect Mesotunnels covered with active functional moieties has been developed in the waste-derived aluminosilicate by the cleavage of the tetrahedrally-coordinated cages, consisted of  −  −   linkages connected via bridging oxygens (), into  −   clusters, where charge neutrality for the negatively-charged non-bridging oxygens () is established by an alkali metal cation. Therefore, higher tunneling is indicative of higher  − 6 7 alkali metal cations and thus higher concentrations of active ion exchange sites, while   − 6 7 metal cation has a passive role of network charge compensator only. Herein, the effect of various reaction parameters on the porosity of the aluminosilicate-based NMF, and thus its ion exchange capability, has been examined.

Effect of reaction temperature In order to investigate the effect of reaction temperature, which has an utmost significance in tuning the porosity of materials, the reaction time and impregnation ratio were retained constant at 3 h and 2 /, respectively. As shown in Table 1, as the reaction temperature increases up to 300˚C, the surface area and pore volume increase. This is related to the fact that the cleavage of the  −  −   into  −  linkages is a very energy-demanding process due to the robustness of the former functional group. Thus, higher temperatures provide higher energy intensity for the rupture of the tetrahedral aluminosilicate cages and tunnel formation, as illustrated in Figure 1, explaining the drastic increase in the surface area and pore volume of the aluminosilicate. Further increase in the reaction temperature results in a sharp decrease in both surface area and pore volume which is ascribed to the distortion of the aluminosilicate network, rather than its

ACS Paragon Plus Environment

Page 8 of 46

Page 9 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

partial bond cleavage, at such high temperatures. (Figure 1) (Table 1) The negatively-charged ions formed on the surface of the aluminosilicate ( −  ) are counterbalanced by two distinct approaches; 1) the cations of the activating agent, 8 9 , form  −  … 8 9 clusters, and 2) when the Al-containing tetrahedral network is ruptured, the :;9 cations, already as the charge compensator for  clusters, migrate to the aluminosilicate surface and take the role of the charge-balancing cation for  −  linkages. Both of these cations can play the role of ion exchange sites to uptake heavy metal ions from aqueous solutions. Therefore, higher surface area is an indication of enhanced bond cleavage and thus higher concentration of ion exchange sites. Figure 2 compares the nitrogen adsorption/desorption isotherms and pore size distribution curves for the aluminosilicate materials functionalized at various reaction temperatures. All the isotherm curves, regardless of the reaction temperature, indicate the type IV isotherms with distinct H3type hysteresis loops, according to the IUPAC classification, characteristic of the mesoporous materials. Despite the existence of a hysteresis loop for all the materials, the hysteresis loop areas are different (see Table S1). FD250 and FD300 exhibit larger hysteresis loop areas, which are translated into larger mesopore surface area, in close agreement with the quantitative data presented in Table 1. (Figure 2) The trend of the change in the pore size of the samples is relatively identical to that of the surface area, where an increase in the reaction temperature from 200˚C to 250˚C also increases the average diameter of the developed pores. Further increase in the reaction temperature to 300˚C

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

results in a slight decrease in the average pore size of the material. This trend is attributed to the parallel occurrence of two phenomena during the functionalization process; the development of micropores and the development and/or enlargement of mesopores as a result of small pore mergence. In the first stage, the micropore development is relatively low and mesopore enlargement is the dominating factor. However, when the temperature is further increased, micropore development dominates the mesopore enlargement validated by a significant increase in the micropore surface area (see Table 1). At an activation temperature of 400˚C, the aluminosilicate network is totally distorted due to the pore collapse or pore contraction 15. Therefore, unlike carbonaceous materials which require pore formation temperatures as high as 800-1000˚C45–49, even slightly high temperatures are detrimental for the pore formation of the siliceous materials, resulting in the recombination of the developed active  −  moieties into undesirable formation of passive  −  −   linkages.

Impregnation ratio Quantity of the activating agent is another key element which significantly affects the textural and surface properties of the aluminosilicate materials and also is a determining factor in the cost effectiveness of the functionalization process. For effective tunneling and functionalization, sufficient amount of activating agent should be available in the reaction medium to cleave the robust  −  −   linkages in the tetrahedral structure of the aluminosilicate. Also, excessive amounts of the activating agent, not only is not economical in terms of the chemical cost and the energy required to heat the activating agent solution, but also may result in “over-cleavage” of the tetrahedral bonds and thus distortion of the structure. In order to evaluate the optimum quantity of the activating agent in the reaction, the effect of the impregnation ratio on the textural

ACS Paragon Plus Environment

Page 10 of 46

Page 11 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

properties of the aluminosilicate has been studied, while the reaction time and temperature are fixed at 3 ℎ and 300˚:, respectively. The activated materials have been denoted as ?@A, where  indicates the impregnation ratio (/) employed in the reaction. As shown in Table 1, when the impregnation ratio is 1 /, incomplete activation occurs resulting in very low surface area and pore volume. This is due to the inadequate amount of the activating agent to break the abundant concentration of the  −  −   linkages. Higher porosity can be achieved by increasing the relative quantity of the activating agent to reach the impregnation ratio of 2 /. This is indicative of the cleavage of more  −  −   linkages which facilitates the creation of mesotunnels and hence higher surface area and pore volume. However, when excessive activating agent amount is introduced into the reaction, an interesting change in the textural properties occurs; the surface area remains unchanged, but the pore volume drastically increases. Supposedly, the surface area and pore volume should increase upon the introduction of large amount of the activating agent due to the formation of more mesotunnels. Nevertheless, at such vigorous basic conditions, a parallel phenomenon concurrently occurs in addition to the tunnel formation, the removal of pore walls separating two adjacent tunnels. This competing phenomenon, schematically illustrated in Figure 3, apparently leads to a decrease in the surface area, while retaining the pore volume unchanged. The coherent add-up of these two phenomena, i.e. tunnel formation and pore wall removal, results in the observed eccentric change in the textural properties of the aluminosilicate. Considering the decrease in the surface area of the aluminosilicate and the increase in the pore volume at such excessive activating agent amount, it can be inferred that the pore wall removal is the prevailing mechanism although the tunneling effect still proceeds in the activation process. The drastic increase in the pore size of the aluminosilicate at excessive amounts of the activating agent also

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

validates the pore wall removal phenomenon. In the case of ion exchange and chemisorption, pore wall removal is not considered as a desired phenomenon, as pore walls accommodate a large fraction of active site which are eliminated upon the removal of the walls. (Figure 3) The nitrogen adsorption-desorption isotherms and pore size distribution curves for aluminosilicates with various impregnation ratios has been presented in Figure 4. The isotherm curve for ?@A1 depicts the poorly-developed mesotunnels, verified by the very small hysteresis loop. Increasing the impregnation ratio to ≥ 2 / clearly increases both the total adsorbed volume of nitrogen into the pores and the hysteresis area, indicating higher pore volume and enhanced mesotunnel formation. Moreover, the pore size distribution of the ?@A4 exhibits a drastic pore widening effect. These isotherm curves are in good agreement with the aforementioned mechanism of mesotunnel formation and pore widening effects. (Figure 4)

Reaction time The effect of the reaction time on the activation intensity of the aluminosilicate has been presented in Table 1. It has been demonstrated that the reaction time up to 5 h does not have a significant effect on the textural properties of the aluminosilicate. Surface area, pore volume and average pore size are approximately identical for the aluminosilicate samples functionalized at reaction times of 1-5 h. Nevertheless, increasing the reaction time to 7 h leads to a dramatic structural collapse of the aluminosilicate, where the surface area and pore volume drastically decrease and pore size increases. This trend of the changes in the porous structure of the aluminosilicate at prolonged reaction times is very similar to that of the excessively high reaction

ACS Paragon Plus Environment

Page 12 of 46

Page 13 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

temperatures. This suggests that the functionalization process takes place in a rapid rate. Therefore, a considerable amount of energy can be saved by minimizing the reaction time to achieve a high ion exchange efficiency. Nitrogen adsorption-desorption curves and pore size distribution diagrams, depicted in Figure 5, also exhibit a drastic decrease in the ultimate pore volume and hysteresis area and a slight inclination towards larger pore sizes when the reaction time exceeds 5 h. (Figure 5)

Optimum reaction conditions Overall, the optimum porous structure and thus ion exchange efficiency is achieved at moderate reaction conditions (reaction temperature: 300˚:, reaction time: 1 ℎ and impregnation ratio: 2 /). Higher reaction times and temperatures do not necessarily lead to any enhancement in the textural properties of the aluminosilicate and surface functional active sites, but may adversely affect the aluminosilicate structural properties and its ion exchange performance. Therefore, the surface of the material functionalized at optimum conditions will be thoroughly characterized to further investigate the developed functional moieties and to elucidate the reaction mechanism. Hereafter, this material will be denoted as A-NMF.

Surface characterization Figure S1 shows the SEM images of NMF and A-NMF. NMF is composed of randomly-oriented glass fibers (mainly calcium aluminosilicate) with carbonaceous materials covering the external surface of the fibers. The fibers exhibit no porous structure. However, after the thermo-alkaline reaction, the fibers are transformed into very porous disordered clusters.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6 displays the FT-IR spectra of the NMF and A-NMF. The band at 470  corresponds to the bending mode of the  −  −   linkages ( species)50. Apparently, the concentration of  −  −   linkages does not significantly vary before and after the activation reaction. This is due to the abundance of the  −  −   linkages both in the bulk and on the surface of the aluminosilicate. Although some  −  −   groups on the aluminosilicate surface are broken into  −   linkages upon functionalization, a major fraction of the  linkages forming the aluminosilicate tetrahedral framework remain intact. Hence, the intensity of the bending mode peak at 470  can be used as reference to compare the intensities of the other peaks altered during the functionalization process. This can give a realistic insight into the concentration of the developed  groups. The broad absorption band in the range of 3600-3200  is attributed to the stretching vibrations of the hydrogen bonded silanol groups and adsorbed water51. Absence of any peak in the 3800-3700  region corresponding to free hydroxyl silanol groups52,53 indicates that no isolated hydroxyl moieties are present on the aluminosilicate surface, implying the high concentration of hydroxyl groups. Also, the band peaked at 1001  is ascribed to the stretching vibrations of the  −  linkages, where  can be either bridging () or non-bridging oxygen ()54. Apparently, the intensities of the peaks at 3471 and 1001  have been significantly increased when compared with the intensity of the reference peak. This illustrates a substantial increase in the concentration of the developed non-bridging  −  sites upon activation. This also verifies the hypothesis of the  −  −   cleavage into  −  linkages and tunnel formation, as discussed earlier. The peak at 1670 , attributed to the hydrogen-bonded hydroxyl group vibrations, intensified upon functionalization, mainly due to the interaction of the water molecules with the abundant silanol groups55. These results are in line with the contact angle measurements for the aluminosilicate samples (see Figure S2), where the NMF exhibits a

ACS Paragon Plus Environment

Page 14 of 46

Page 15 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

hydrophobic behavior with a contact angle of 120˚, while the functionalized A-NMF is highly hydrophilic. This hydrophilicity is established by the development of water-friendly silanol moieties, making it a suitable candidate to be used in aqueous media. (Figure 6) X-ray photoelectron spectroscopy (XPS), as a powerful analysis tool for accurately investigating the atomic composition and environment on the surface layer, has been employed to further investigate the surface functional moieties developed on the A-NMF upon activation (see Figure 7). Substantial differences in the surface chemical composition of the original and functionalized materials were observed in the XPS survey spectra. The appearance of a peak at 292 eV, attributed to the K 2p, for A-NMF validates the creation of potential ion exchange active sites on the material surface56. The intensity of the C 1s peak at 285 eV significantly decreased due to the conversion of carbon impurities to CO2 at high temperatures during the activation process. Also, the intensity of the O 1s peak at around 530 eV has been considerably increased upon the functionalization process which is related to the transformation of the  −  −   into two  −  linkages as well as the exposure of higher amounts of  to the surface because of the tunneling effect. The high resolution spectra for the O 1s peak, as illustrated in the inset of Figure 7, reveals the considerable shift of the O 1s peak to a lower binding energy (from 533.1 eV to 530.8 eV) after the functionalization process. This shift can be explained by the proposed transformation mechanism of  into  as a result of the hydroxylation reaction57. The functionalization process results in the cleavage of the  −  −   (where O is a bridging oxygen) into  −  …  (where O is a non-bridging oxygen and  = 8 9 :;9 ). Thus, the more electropositive 8 9 or :;9 will increase the electron density on the , resulting in the significant shift of the oxygen peak toward lower binding energy57.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(Figure 7)

Ion exchange efficiency The aim of the functionalization process is to increase the concentration of the  species and thus active ion exchange sites. The ion exchange efficiency of the functionalized aluminosilicate material (A-NMF) has been evaluated for several heavy metals and tabulated in Table 2. The heavy metal uptake capacities of the A-NMF for all the metals are much higher than those reported in other literature for other siliceous materials, including zeolites and silica. This exceptionally higher capacity is attributed to the fact that the distinction between the alkali metal cations functioning as charge compensator and network modifier is often disregarded in most literature. As addressed earlier, the alkali metal cations as charge compensators render charge neutrality for the   clusters. In order to retain the electroneutrality in the aluminosilicate network, these alkali metal cations can only locally move along the network. Hence, their diffusion to the aluminosilicate surface is obstructed by the high electrostatic interaction between the   clusters and the cations58, and as a result, the feasibility of ion exchange with these charge balancing cations is low. In this study, the concentration of the network modifying alkali metal cations bound to the  species is significantly increased upon functionalization and some of the calcium cations, originally as charge balancing cations, migrate to the surface in the course of tetrahedral network cleavage and act as active ion exchange sites. The increase in the concentration of available ion exchange sites on the functionalized aluminosilicate material accounts for its ultra-high heavy metal uptake. (Table 2)59–85

ACS Paragon Plus Environment

Page 16 of 46

Page 17 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

In order to evaluate the competitiveness of the heavy metal uptake performance of this wastederived functionalized aluminosilicate with the commercial ion exchange resins, the efficiencies of several well-known commercial resins under the same experimental conditions, i.e. heavy metal concentration, pH value, ion exchange dosage and temperature, have been examined. According to Table 2, it has been shown that the ion exchange capacity of the waste-based A-NMF is also higher than the commercial resins under similar conditions. The high ion exchange performance of A-NMF together with the facile inexpensive functionalization technique enhances the feasibility of its commercialization. The attractiveness of this innovative approach lies on the dual environmental benefit, i.e. sustainable value-added upcycling of a problematic waste initially destined to landfill and its utilization for heavy metal-laden wastewater treatment. In addition to the high efficiency of the ion exchange material produced at optimum conditions, its cost is also expected to be low due to several reasons; the initial precursor is a wastederived material and thus has low value; the yield of the material is as high as 70-80 wt%; the activation temperature is as low as 250℃; and the activation process is a simple practical one. All these factors lead to the cost-effectiveness of the resultant ion exchange material for industrialization purposes. However, it is of note that the functional moieties developed on ANMF is very sensitive to humidity and temperature and thus, should be kept at controlled environments to retain its high efficiency.

Conclusion A very efficient ion exchange material has been sustainably developed by upcycling a wastederived aluminosilicate material (NMF) as precursor. The hydrophobic nonporous NMF has been

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

transformed into functionalized hydrophilic ion exchange material via the cleavage of passive

−  −  bonds into reactive −  −  linkages (where  and  denote bridging and non-bridging oxygens, respectively). This bond cleavage takes place using a thermo-alkaline reaction, where tunnels covered with functional moieties are developed on the aluminosilicate surface by tetrahedral structure rupture. In this study, the effects of several reaction variables, such as the amount of the activating agent, reaction time and reaction temperature, on the creation of these tunnels and surface functional groups have been examined. It has been found that when the reaction time and temperature exceeds 5 h and 300˚C, respectively, the aluminosilicate structure partially collapses, leading to a drastic decrease in the surface area. Also, when the amount of the activating agent employed in the reaction is not sufficient (impregnation ratio < 2 /), the tunneling effect is not properly achieved. The heavy metal uptake capacity of the functionalized aluminosilicate under the optimized reaction conditions has been shown to be exceptionally superior to those of either synthesized or natural zeolites and also higher than the commercial ion exchange resins. This superiority is related to the high concentration of surface ion exchange sites achieved by the functionalization process. The competitive ion exchange capacity of this material combined with the cost-effective functionalization technique employed to upcycle this waste siliceous material into highperformance ion exchange material makes it a suitable material for heavy metal-laden wastewater treatment.

ASSOCIATED CONTENT Supporting Information. SEM graphs, contact angle measurements, and hysteresis data. The information is available free of charge via the ACS Publication website at http://pubs.acs.org.

ACS Paragon Plus Environment

Page 18 of 46

Page 19 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

AUTHOR INFORMATION Corresponding Authors. *Phone: +852-23588412, Fax: +852-23580054, E-mail: [email protected]. Notes. The authors declare no competing financial interests.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 46

References

(1)

Choi, M.; Cho, H. S.; Srivastava, R.; Venkatesan, C.; Choi, D.-H.; Ryoo, R. Amphiphilic organosilane-directed synthesis of crystalline zeolite with tunable mesoporosity. Nat. Mater. 2006, 5, 718–723.

(2)

Morris, R. E.; Čejka, J. Exploiting chemically selective weakness in solids as a route to new porous materials. Nat. Chem. 2015, 7, 381–388.

(3)

Doménech, A.; García, H.; Marquet, J.; Bourdelande, J. L.; Herance, J. R. Modelling electrocatalysis of hydroquinone oxidation by nicotinamide adenine dinucleaotide coenzyme encapsulated within SBA-15 and MCM-41 mesoporous aluminosilicates. Electrochim. Acta 2006, 51, 4897–4908.

(4)

Wang, L.; Xiao, F.-S. Nanoporous catalysts for biomass conversion. Green Chem. 2014, 17, 24–39.

(5)

Li, T.; Liu, H.; Fan, Y.; Yuan, P.; Shi, G.; Bi, X. T.; Bao, X. Synthesis of zeolite Y from natural aluminosilicate minerals for fluid catalytic cracking application. Green Chem. 2012, 14, 3255–3459.

(6)

Mo, N.; Savage, P. E. Hydrothermal Catalytic Cracking of Fatty Acids with HZSM-5. ACS Sustain. Chem. Eng. 2014, 2, 88–94.

(7)

Jeganathan, M.; Pitchumani, K. Solvent-Free Syntheses of 1,5-Benzodiazepines Using HY Zeolite as a Green Solid Acid Catalyst. ACS Sustain. Chem. Eng. 2013, 1, 316−324.

(8)

Wu, C.; Dong, L.; Onwudili, J.; Williams, P. T.; Huang, J. Effect of Ni particle location within the mesoporous MCM-41 support for hydrogen production from the catalytic gasification of biomass. ACS Sustain. Chem. Eng. 2013, 1, 1083–1091.

(9)

Zhang, W.; Chen, J.; Liu, R.; Wang, S.; Chen, L.; Li, K.; Accepted, J. Hydrodeoxygenation of lignin-derived phenolic monomers and dimers to alkane fuels over bifunctional zeolite-supported metal catalysts. Substainable Chem. Eng. 2014, 2, 683−691.

(10)

Phair, J. W. Green chemistry for sustainable cement production and use. Green Chem. 2006, 8, 763–780.

ACS Paragon Plus Environment

Page 21 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(11)

Faramawy, S.; El-Naggar, a. Y.; El-Fadly, a. M.; El-Sabagh, S. M.; Ibrahim, a. a. Silica, alumina and aluminosilicates as solid stationary phases in gas chromatography. Arab. J. Chem. 2011.

(12)

Wei, X.; Wang, W.; Xiao, J.; Zhang, L.; Chen, H.; Ding, J. Hierarchically porous aluminosilicates as the water vapor adsorbents for dehumidification. Chem. Eng. J. 2013, 228, 1133–1139.

(13)

Majano, G.; Ng, E.-P.; Lakiss, L.; Mintova, S. Nanosized molecular sieves utilized as an environmentally friendly alternative to antioxidants for lubricant oils. Green Chem. 2011, 13, 2435–2440.

(14)

Skorina, T.; Allanore, A. Aqueous alteration of potassium-bearing aluminosilicate minerals: from mechanism to processing. Green Chem. 2015, 17, 2123–2136.

(15)

Huang, Z.; Miao, H.; Li, J.; Wei, J.; Kawi, S.; Lai, M. W. Modifier-enhanced supercritical CO2 extraction of organic template from aluminosilicate MCM-41 materials: Effect of matrix Al/Si ratios and different modifiers. Sep. Purif. Technol. 2013, 118, 170–178.

(16)

Hadi, P.; Gao, P.; Barford, J. P.; McKay, G. Novel application of the nonmetallic fraction of the recycled printed circuit boards as a toxic heavy metal adsorbent. J. Hazard. Mater. 2013, 252-253, 166–170.

(17)

Detoni, C.; Gierlich, C. H.; Rose, M.; Palkovits, R. Selective Liquid Phase Adsorption of 5-Hydroxymethylfurfural on Nanoporous Hyper-Cross-Linked Polymers. ACS Sustain. Chem. Eng. 2014, 2, 2407−2415.

(18)

Ahmed, Z. T.; Hand, D. W. Direct Adsorption Isotherms of AEAs and Fly Ash: α-Olefin Sulfonate and Combination Admixtures. ACS Sustain. Chem. Eng. 2015, 3, 216–223.

(19)

Yan, F.; Jiang, J.; Li, K.; Tian, S.; Zhao, M.; Chen, X. Performance of Coal Fly Ash Stabilized, CaO-based Sorbents under Different Carbonation–Calcination Conditions. ACS Sustain. Chem. Eng. 2015, 3, 2092–2099.

(20)

Jeganathan, M.; Dhakshinamoorthy, A.; Pitchumani, K. One-pot synthesis of propargylamines using Ag(I)-exchanged K10 montmorillonite clay as reusable catalyst in water. ACS Sustain. Chem. Eng. 2014, 2, 781–787.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 46

(21)

Kang, D.-Y.; Brunelli, N. A.; Yucelen, G. I.; Venkatasubramanian, A.; Zang, J.; Leisen, J.; Hesketh, P. J.; Jones, C. W.; Nair, S. Direct synthesis of single-walled aminoaluminosilicate nanotubes with enhanced molecular adsorption selectivity. Nat. Commun. 2014, 5, 3342.

(22)

Farnan, I. Structural chemistry: Oxygen bridges in molten glass. Nature 1997, 390, 14–15.

(23)

Stebbins, J. F.; Xu, Z. NMR evidence for excess non-bridging oxygen in an aluminosilicate glass. Nature 1997, 390, 60–62.

(24)

O’Connor, S. J.; MacKenzie, K. J. D.; Smith, M. E.; Hanna, J. V. Ion exchange in the charge-balancing sites of aluminosilicate inorganic polymers. J. Mater. Chem. 2010, 20, 10234–10240.

(25)

Perić, J.; Trgo, M.; Vukojević Medvidović, N. Removal of zinc, copper and lead by natural zeolite-a comparison of adsorption isotherms. Water Res. 2004, 38, 1893–1899.

(26)

Erdem, E.; Karapinar, N.; Donat, R. The removal of heavy metal cations by natural zeolites. J. Colloid Interface Sci. 2004, 280, 309–314.

(27)

Sprynskyy, M.; Buszewski, B.; Terzyk, A. P.; Namieśnik, J. Study of the selection mechanism of heavy metal (Pb2+, Cu2+, Ni2+, and Cd2+) adsorption on clinoptilolite. J. Colloid Interface Sci. 2006, 304, 21–28.

(28)

Alkan, M.; Kalay, B.; Doğan, M.; Demirbaş, O. Removal of copper ions from aqueous solutions by kaolinite and batch design. J. Hazard. Mater. 2008, 153, 867–876.

(29)

Yavuz, O.; Altunkaynak, Y.; Güzel, F. Removal of copper, nickel, cobalt and manganese from aqueous solution by kaolinite. Water Res. 2003, 37, 948–952.

(30)

Yang, S.; Sheng, G.; Tan, X.; Hu, J.; Du, J.; Montavon, G.; Wang, X. Determination of Ni(II) uptake mechanisms on mordenite surfaces: A combined macroscopic and microscopic approach. Geochim. Cosmochim. Acta 2011, 75, 6520–6534.

(31)

Kodama, T.; Komarneni, S. Na-4-mica: Cd2+, Ni2+, Co2+, Mn2+ and Zn2+ ion exchange. J. Mater. Chem. 1999, 9, 533–539.

(32)

Sen Gupta, S.; Bhattacharyya, K. G. Adsorption of metal ions by clays and inorganic solids. RSC Adv. 2014, 4, 28537.

ACS Paragon Plus Environment

Page 23 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(33)

Bouyer, F.; Geneste, G.; Ispas, S.; Kob, W.; Ganster, P. Water solubility in calcium aluminosilicate glasses investigated by first principles techniques. J. Solid State Chem. 2010, 183, 2786–2796.

(34)

Huang, K.; Guo, J.; Xu, Z. Recycling of waste printed circuit boards: a review of current technologies and treatment status in China. J. Hazard. Mater. 2009, 164, 399–408.

(35)

Li, J.; Lu, H.; Guo, J.; Xu, Z.; Zhou, Y. Recycle technology for recovering resources and products from waste printed circuit boards. Environ. Sci. Technol. 2007, 41, 1995–2000.

(36)

Hadi, P.; Xu, M.; Hui, C.-W.; Lin, C. S. K.; McKay, G. Waste Printed Circuit Board Recycling Techniques and Product Utilization. J. Hazard. Mater. 2015, 283, 234–243.

(37)

Jang, Y.-C.; Townsend, T. G. Leaching of lead from computer printed wire boards and cathode ray tubes by municipal solid waste landfill leachates. Environ. Sci. Technol. 2003, 37, 4778–4784.

(38)

Hadi, P.; Ning, C.; Ouyang, W.; Xu, M.; Lin, C. S. K.; McKay, G. Toward environmentally-benign utilization of nonmetallic fraction of waste printed circuit boards as modifier and precursor. Waste Manag. 2015, 35C, 236–246.

(39)

Duan, C.; Wen, X.; Shi, C.; Zhao, Y.; Wen, B.; He, Y. Recovery of metals from waste printed circuit boards by a mechanical method using a water medium. J. Hazard. Mater. 2009, 166, 478–482.

(40)

Petter, P. M. H.; Veit, H. M.; Bernardes, A. M. Evaluation of gold and silver leaching from printed circuit board of cellphones. Waste Manag. 2014, 34, 475–482.

(41)

Hadi, P.; Barford, J.; McKay, G. Toxic heavy metal capture using a novel electronic wastebased material-mechanism, modeling and comparison. Environ. Sci. Technol. 2013, 47, 8248–8255.

(42)

Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1936, 407, 309–313.

(43)

Shull, C. G. The Determination of Pore Size Distribution from Gas Adsorption Data. J. Am. Chem. Soc. 1948, 70, 1405–1410.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 46

(44)

Barrett, E. P.; Joyner, L. G.; Halenda, P. P. The Determination of Pore Volume and Area Distributions in Porous Substances. I. Computations from Nitrogen Isotherms. J. Am. Chem. Soc. 1951, 73, 373–380.

(45)

Sangchoom, W.; Mokaya, R. Valorization of Lignin Waste: Carbons from Hydrothermal Carbonization of Renewable Lignin as Superior Sorbents for CO2 and Hydrogen Storage. ACS Sustain. Chem. Eng. 2015, 3, 1658–1667.

(46)

Hadi, P.; Xu, M.; Ning, C.; Lin, C.; McKay, G. A critical review on preparation, characterization and utilization of sludge-derived activated carbons for wastewater treatment. Chem. Eng. J. 2015, 260, 895–906.

(47)

Hu, S.; Hsieh, Y. Preparation of Activated Carbon and Silica Particles from Rice Straw. ACS Sustain. Chem. Eng. 2014, 2, 726–734.

(48)

Shi, Y.; Zhang, X.; Liu, G. Activated Carbons Derived from Hydrothermally Carbonized Sucrose: Remarkable Adsorbents for Adsorptive Desulfurization. ACS Sustain. Chem. Eng. 2015, 3, 2237–2246.

(49)

Choma, J.; Marszewski, M.; Osuchowski, L.; Jagiello, J.; Dziura, A.; Jaroniec, M. Adsorption Properties of Activated Carbons Prepared from Waste CDs and DVDs. ACS Sustain. Chem. Eng. 2015, 3, 733–742.

(50)

Betiha, M. A.; Menoufy, M. F.; Al-Sabagh, A. M.; Hassan, H. M. A.; Mahmoud, S. A. Acidic mesostructured aluminosilicates assembled from economic acidic template characterized by catalytic cracking reactions. Microporous Mesoporous Mater. 2015, 204, 15–24.

(51)

Dugas, V.; Chevalier, Y. Surface hydroxylation and silane grafting on fumed and thermal silica. J. Colloid Interface Sci. 2003, 264, 354–361.

(52)

Neelakandeswari, N.; Sangami, G.; Emayavaramban, P.; Ganesh Babu, S.; Karvembu, R.; Dharmaraj, N. Preparation and characterization of nickel aluminosilicate nanocomposites for transfer hydrogenation of carbonyl compounds. J. Mol. Catal. A Chem. 2012, 356, 90– 99.

(53)

Garbassi, F.; Balducci, L.; Chiurlo, P.; Deiana, L. A study of surface modification of silica using XPS, DRIFT and NMR. Appl. Surf. Sci. 1995, 84, 145–151.

ACS Paragon Plus Environment

Page 25 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(54)

Zhou, C.; Sun, T.; Gao, Q.; Alshameri, A.; Zhu, P.; Wang, H.; Qiu, X.; Ma, Y.; Yan, C. Synthesis and characterization of ordered mesoporous aluminosilicate molecular sieve from natural halloysite. J. Taiwan Inst. Chem. Eng. 2014, 45, 1073–1079.

(55)

Borodina, E.; Karpov, S. I.; Selemenev, V. F.; Schwieger, W.; Maracke, S.; Fröba, M.; Rößner, F. Surface and texture properties of mesoporous silica materials modified by silicon-organic compounds containing quaternary amino groups for their application in base-catalyzed reactions. Microporous Mesoporous Mater. 2015, 203, 224–231.

(56)

Sawyer, R.; Nesbitt, H. W.; Secco, R. a. High resolution X-ray Photoelectron Spectroscopy (XPS) study of K2O–SiO2 glasses: Evidence for three types of O and at least two types of Si. J. Non. Cryst. Solids 2012, 358, 290–302.

(57)

Miura, Y.; Matsumoto, S.; Nanba, T.; Akazawa, T. X-ray photoelectron spectroscopy of sodium aluminosilicate glasses. Phys. Chem. Glas. 2000, 41, 24–31.

(58)

Iversen, B. B.; Latturner, S.; Stucky, G. The Nature of Chemical Interactions in Anhydrous Sodalites, M6(AlSiO4)6, M= Na, Ag, Tl. Chem. Mater. 1999, 11, 2912–2918.

(59)

Kocaoba, S.; Orhan, Y.; Akyüz, T. Kinetics and equilibrium studies of heavy metal ions removalby use of natural zeolite. Desalination 2007, 214, 1–10.

(60)

Oter, O.; Akcay, H. Use of natural clinoptilolite to improve water quality: sorption and selectivity studies of lead(II), copper(II), zinc(II), and nickel(II). Water Environ. Res. 2007, 79, 329–335.

(61)

Li, C.; Zhong, H.; Wang, S.; Xue, J.; Zhang, Z. A novel conversion process for waste residue: Synthesis of zeolite from electrolytic manganese residue and its application to the removal of heavy metals. Colloids Surfaces A Physicochem. Eng. Asp. 2015, 470, 258– 267.

(62)

El-Mekkawi, D. M.; Selim, M. M. Removal of Pb2+ from water by using Na-Y zeolites prepared from Egyptian kaolins collected from different sources. J. Environ. Chem. Eng. 2014, 2, 723–730.

(63)

El-Kamash, A. M.; Zaki, A. A.; El Geleel, M. A. Modeling batch kinetics and thermodynamics of zinc and cadmium ions removal from waste solutions using synthetic zeolite A. J. Hazard. Mater. 2005, 127, 211–220.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 46

(64)

Wan Ngah, W. S.; Teong, L. C.; Toh, R. H.; Hanafiah, M. a K. M. Comparative study on adsorption and desorption of Cu(II) ions by three types of chitosan-zeolite composites. Chem. Eng. J. 2013, 223, 231–238.

(65)

Zhang, Y.; Yan, W.; Sun, Z.; Pan, C.; Mi, X.; Zhao, G.; Gao, J. Fabrication of porous zeolite/chitosan monoliths and their applications for drug release and metal ions adsorption. Carbohydr. Polym. 2015, 117, 657–665.

(66)

Anirudhan, T. S.; Suchithra, P. S. Heavy metals uptake from aqueous solutions and industrial wastewaters by humic acid-immobilized polymer/bentonite composite: Kinetics and equilibrium modeling. Chem. Eng. J. 2010, 156, 146–156.

(67)

Jha, V. K.; Matsuda, M.; Miyake, M. Sorption properties of the activated carbon-zeolite composite prepared from coal fly ash for Ni(2+), Cu(2+), Cd(2+) and Pb(2+). J. Hazard. Mater. 2008, 160, 148–153.

(68)

Mureseanu, M.; Reiss, A.; Stefanescu, I.; David, E.; Parvulescu, V.; Renard, G.; Hulea, V. Modified SBA-15 mesoporous silica for heavy metal ions remediation. Chemosphere 2008, 73, 1499–1504.

(69)

Rad, L. R.; Momeni, A.; Ghazani, B. F.; Irani, M.; Mahmoudi, M.; Noghreh, B. Removal of Ni2+ and Cd2+ ions from aqueous solutions using electrospun PVA/zeolite nanofibrous adsorbent. Chem. Eng. J. 2014, 256, 119–127.

(70)

Qiu, W.; Zheng, Y. Removal of lead, copper, nickel, cobalt, and zinc from water by a cancrinite-type zeolite synthesized from fly ash. Chem. Eng. J. 2009, 145, 483–488.

(71)

Qi, G.; Lei, X.; Li, L.; Yuan, C.; Sun, Y.; Chen, J.; Chen, J.; Wang, Y.; Hao, J. Preparation and evaluation of a mesoporous calcium-silicate material (MCSM) from coal fly ash for removal of Co(II) from wastewater. Chem. Eng. J. 2015, 279, 777–787.

(72)

Nascimento, M.; Soares, P. S. M.; Souza, V. P. De. Adsorption of heavy metal cations using coal fly ash modified by hydrothermal method. Fuel 2009, 88, 1714–1719.

(73)

Wang, P.; Du, M.; Zhu, H.; Bao, S.; Yang, T.; Zou, M. Structure regulation of silica nanotubes and their adsorption behaviors for heavy metal ions: pH effect, kinetics, isotherms and mechanism. J. Hazard. Mater. 2015, 286, 533–544.

(74)

El-Toni, A. M.; Habila, M. a.; Ibrahim, M. A.; Labis, J. P.; ALOthman, Z. a. Simple and

ACS Paragon Plus Environment

Page 27 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

facile synthesis of amino functionalized hollow core-mesoporous shell silica spheres using anionic surfactant for Pb(II), Cd(II), and Zn(II) adsorption and recovery. Chem. Eng. J. 2014, 251, 441–451. (75)

Lin, X. Z.; Ma, T. Y.; Yuan, Z. Y. Titania-silica-phosphonate triconstituent hybrid mesoporous materials as adsorbents in gas and liquid phases. Chem. Eng. J. 2011, 166, 1144–1151.

(76)

Al-Harahsheh, M. S.; Al Zboon, K.; Al-Makhadmeh, L.; Hararah, M.; Mahasneh, M. Fly ash based geopolymer for heavy metal removal: A case study on copper removal. J. Environ. Chem. Eng. 2015, 3, 1669–1677.

(77)

Visa, M.; Chelaru, A. M. Hydrothermally modified fly ash for heavy metals and dyes removal in advanced wastewater treatment. Appl. Surf. Sci. 2014, 303, 14–22.

(78)

Dimitrova, S. V. Metal sorption on blast-furnace slag. Water Res. 1996, 30, 228–232.

(79)

Álvarez-Ayuso, E.; García-Sánchez, a.; Querol, X. Purification of metal electroplating waste waters using zeolites. Water Res. 2003, 37, 4855–4862.

(80)

Hui, K. S.; Chao, C. Y. H.; Kot, S. C. Removal of mixed heavy metal ions in wastewater by zeolite 4A and residual products from recycled coal fly ash. J. Hazard. Mater. 2005, 127, 89–101.

(81)

Vengris, T.; Binkiene, R.; Sveikauskaite, a. Nickel, copper and zinc removal from waste water by a modified clay sorbent. Appl. Clay Sci. 2001, 18, 183–190.

(82)

Pehlivan, E.; Altun, T. The study of various parameters affecting the ion exchange of Cu2+, Zn2+, Ni2+, Cd2+, and Pb2+ from aqueous solution on Dowex 50W synthetic resin. J. Hazard. Mater. 2006, 134, 149–156.

(83)

Demirbas, A.; Pehlivan, E.; Gode, F.; Altun, T.; Arslan, G. Adsorption of Cu(II), Zn(II), Ni(II), Pb(II), and Cd(II) from aqueous solution on Amberlite IR-120 synthetic resin. J. Colloid Interface Sci. 2005, 282, 20–25.

(84)

Rengaraj, S.; Yeon, J. W.; Kim, Y.; Jung, Y.; Ha, Y. K.; Kim, W. H. Adsorption characteristics of Cu(II) onto ion exchange resins 252H and 1500H: Kinetics, isotherms and error analysis. J. Hazard. Mater. 2007, 143, 469–477.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(85)

Pehlivan, E.; Altun, T. Ion-exchange of Pb2+, Cu2+, Zn2+, Cd2+, and Ni2+ ions from aqueous solution by Lewatit CNP 80. J. Hazard. Mater. 2007, 140, 299–307.

ACS Paragon Plus Environment

Page 28 of 46

Page 29 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Table 1. Textural properties of the materials functionalized at different reaction conditions. Sample ID

FGHI

FJNO

PQ

ST

(JK . LM)

(JK . LM)

(OJR . LM )

(UJ)

Effect of reaction temperature (impregnation ratio= K V/V and reaction time = R W) FD200

317

0

0.615

7.8

FD250

358

4

0.832

9.3

FD300

383

18

0.851

8.9

FD400

282

0

0.586

8.3

Effect of impregnation ratio (reaction temperature= RXX℃ and reaction time = R W) IMP1

143

0

0.302

8.4

IMP2

383

18

0.851

8.9

IMP4

331

0

1.029

12.4

Effect of reaction time (reaction temperature= RXX℃ and impregnation ratio= K V/V) d1h

368

4

0.848

9.2

d3h

383

18

0.851

8.9

d5h

378

10

0.853

9.0

d7h

262

7

0.678

10.3

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Page 30 of 46

Table 2. Comparison of the ion exchange capacity of the A-NMF with those in the literature and the commercial ion exchange resins.

Maximum ion exchange (JJYZ. LM) Material

Ref.

Cu2+

Pb2+

Zn2+

Co2+

Ni2+

Cd2+

NMF

< 0.1

< 0.1

< 0.1

< 0.1

< 0.1

< 0.1

This work

A-NMF

2.82

3.21

2.00

3.12

3.11

2.10

This work

1.67

1.11

59

Siliceous Material Heavy Metal Uptake Capacities Natural Zeolite

1.18

Natural Clinoptilolite

0.84

0.48

0.19

EMR-derived Zeolite

0.13

60

2.19

61 62

1.50

Na-Y Zeolite

1.85

Synthetic Zeolite A Chitosan-zeolite Composites

0.81

Zeolite/Chitosan Monoliths

1.34

Acid-Immobilized polymer/bentonite

0.74

Activated carbon-Zeolite Composite

1.72

Modified SBA-15

0.92

63 64 65

0.29 0.68

66

0.58

2.65

1.20 0.4

ACS Paragon Plus Environment

0.32

0.36

1.44

67 68

Page 31 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

ACS Sustainable Chemistry & Engineering

1.04

Electrospun PVA/Zeolite nanofiber 2.08

Cancrinite type Zeolite

2.13

1.15

1.24

0.77

70

1.53

71

2.63

Modified Coal Fly Ash Hydrothermal treated Coal Fly Ash

0.95

0.69

Functionalized Silica Nanotubes

0.49

Amino-functionalized Silica

0.94

Ti-Si-P Hybrid Material

0.45

69

72

0.62 1.77 2.95

0.89

73

1.69

74

0.45

75 76

Fly Ash Based Geopolymer

0.91

Hydrothermally Modified Fly Ash

0.89

Blast-furnace Slag

2.10

1.58

0.95

Synthetic Zeolite

0.80

0.50

0.34

Zeolite 4A

0.63

0.62

Modified Clay

0.40

0.23

0.78

0.29

77 78

0.45

79

0.20

80

0.58

81

Commercial Ion Exchanger Heavy Metal Uptake Capacities Dowex 50W

0.35

0.23

0.30

0.12

0.25

82

Amberlite IR-120

0.34

0.41

1.3

0.82

0.90

83

Amberjet 1500H

0.39

84

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Page 32 of 46

0.32

0.04

85

1.97

1.78

1.21

This work

1.70

2.10

1.88

1.69

This work

0.41

0.71

0.67

0.30

This work

Lewatit CNP 80

0.16

0.35

0.31

Lewatit TP207

1.98

1.78

1.63

Suqing D401

2.00

1.88

MCM-41

0.71

0.68

ACS Paragon Plus Environment

Page 33 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure Legends Figure 1 ¦ Schematic representation of I − [ − I bond cleavage and tunneling effect via the thermo-alkaline reaction (Silicon, black; Aluminum, purple; oxygen, red; calcium, green; potassium, blue; hydrogen yellow). This figure illustrates the hypothesized functionalization reaction in which the passive robust  −  −   linkages are cleaved in the course of thermo-alkaline reaction, resulting in the formation of ion exchange active sites. Some of the calcium ions, originally as charge balancing cations for   clusters, migrate to the surface acting as network modifying cations.

Figure 2 ¦ (a) Nitrogen adsorption-desorption isotherm, and (b) pore size distribution for the samples prepared at different activation temperature. The effect of the thermo-alkaline reaction temperature on the textural properties of the materials depicts that all the prepared materials, irrespective of the reaction temperature, show a type IV isotherm curve with a distinct hysteresis. However, At moderate reaction temperatures (FD250 and FD300), the pore volume and surface area have the highest values.

Figure 3 ¦ Schematic representation of pore wall removal caused by aggressive reaction conditions (Silicon, black; Aluminum, purple; oxygen, red). The

33

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

hypothesized removal of pore walls leads to a decrease in the surface area, while keeping the pore volume unchanged.

Figure 4 ¦ (a) Nitrogen adsorption desorption isotherm, and (b) pore size distribution for the samples prepared at different impregnation ratios. When the amount of the activating agent is not sufficient (impregnation ratio of 1 /), the pore formation is not well-developed. Also, very high amount of the activating agent in the reaction medium results in the removal of pore walls, thus drastically increasing the average pore size.

Figure 5 ¦ (a) Nitrogen adsorption desorption isotherm, and (b) pore size distribution for the samples prepared at different reaction times. This verifies that the thermo-alkaline reaction is a very fast process and very prolonged reaction time has an adverse effect on the textural properties of the material.

Figure 6 ¦ FTIR spectra for the precursor and functionalized aluminosilicate. The functionalization reaction results in a significant increase in the concentration of the NBO moieties, which are counter-balanced with alkali metal cations. These cations can play the role of ion exchange sites on the aluminosilicate.

34

ACS Paragon Plus Environment

Page 34 of 46

Page 35 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 7 ¦ XPS spectra for the precursor and functionalized aluminosilicate. The appearance of the K 2p peak on the functionalized aluminosilicate material and the increase in the intensities of the Ca 2p and O 1s peaks are in line with the hypothesized reaction mechanism, discussed earlier. A considerable shift of the O 1s peak to a lower binding energy (as depicted in the inset Figure) upon the functionalization process suggests the increase in the electron density of the oxygen as a result of the transformation mechanism of  into  linkages.

35

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Thermo-alkaline reaction

Figure 1 ¦ Schematic representation of I − [ − I bond cleavage and tunneling effect via the thermo-alkaline reaction (Silicon, black; Aluminum, purple; oxygen, red; calcium, green; potassium, blue; hydrogen yellow). This figure illustrates the hypothesized functionalization reaction in which the passive robust  −  −   linkages are cleaved in the course of thermo-alkaline reaction, resulting in the formation of ion exchange active sites. Some of the calcium ions, originally as charge balancing cations for   clusters, migrate to the surface acting as network modifying cations.

36

ACS Paragon Plus Environment

Page 36 of 46

Page 37 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 2 ¦ (a) Nitrogen adsorption-desorption isotherm, and (b) pore size distribution for the samples prepared at different activation temperature. The effect of the thermo-alkaline reaction temperature on the textural properties of the materials depicts that all the prepared materials, irrespective of the reaction temperature, show a type IV isotherm curve with a distinct hysteresis. However, At moderate reaction temperatures (FD250 and FD300), the pore volume and surface area have the highest values.

37

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Pore wall removal

Figure 3 ¦ Schematic representation of pore wall removal caused by aggressive reaction conditions (Silicon, black; Aluminum, purple; oxygen, red). The hypothesized removal of pore walls leads to a decrease in the surface area, while keeping the pore volume unchanged.

38

ACS Paragon Plus Environment

Page 38 of 46

Page 39 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 4 ¦ (a) Nitrogen adsorption desorption isotherm, and (b) pore size distribution for the samples prepared at different impregnation ratios. When the amount of the activating agent is not sufficient (impregnation ratio of 1 /), the pore formation is not well-developed. Also, very high amount of the activating agent in the reaction medium results in the removal of pore walls, thus drastically increasing the average pore size.

39

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5 ¦ (a) Nitrogen adsorption desorption isotherm, and (b) pore size distribution for the samples prepared at different reaction times. This verifies that the thermo-alkaline reaction is a very fast process and very prolonged reaction time has an adverse effect on the textural properties of the material.

40

ACS Paragon Plus Environment

Page 40 of 46

Page 41 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 6 ¦ FTIR spectra for the precursor and functionalized aluminosilicate. The functionalization reaction results in a significant increase in the concentration of the NBO moieties, which are counter-balanced with alkali metal cations. These cations can play the role of ion exchange sites on the aluminosilicate.

41

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 7 ¦ XPS spectra for the precursor and functionalized aluminosilicate. The appearance of the K 2p peak on the functionalized aluminosilicate material and the increase in the intensities of the Ca 2p and O 1s peaks are in line with the hypothesized reaction mechanism, discussed earlier. A considerable shift of the O 1s peak to a lower binding energy (as depicted in the inset Figure) upon the functionalization process suggests the increase in the electron density of the oxygen as a result of the transformation mechanism of  into  linkages.

42

ACS Paragon Plus Environment

Page 42 of 46

Page 43 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Valorization of an electronic waste-derived aluminosilicate – Surface functionalization and porous structure tuning sc-2015-01523e

Chao Ninga, Pejman Hadia, Carol Sze Ki Linb and Gordon McKaya,c* a

Chemical and Biomolecular Engineering Department, Hong Kong University of Science and

Technology, Clear Water Bay Road, Hong Kong SAR b

School of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Hong

Kong SAR c

Division of Sustainable Development, College of Science, Engineering and Technology,

Hamad Bin Khalifa University, Qatar Foundation, Doha, Qatar *

Tel: +852 23588412, Fax: +852 23580054, E-mail: [email protected]

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 46

Network modifier ions as ion exchange sites

Developed mesotunnels

Thermo-alkali reaction

Inaccessible charge compensator ions

Si

Al

O

Ca

K

ACS Paragon Plus Environment

H

Page 45 of 46

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Valorization of an electronic waste-derived aluminosilicate – Surface functionalization and porous structure tuning sc-2015-01523e

Chao Ninga, Pejman Hadia, Carol Sze Ki Linb and Gordon McKaya,c* a

Chemical and Biomolecular Engineering Department, Hong Kong University of Science and

Technology, Clear Water Bay Road, Hong Kong SAR b

School of Energy and Environment, City University of Hong Kong, Tat Chee Avenue, Hong

Kong SAR c

Division of Sustainable Development, College of Science, Engineering and Technology,

Hamad Bin Khalifa University, Qatar Foundation, Doha, Qatar *

Tel: +852 23588412, Fax: +852 23580054, E-mail: [email protected]

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 46

Network modifier ions as ion exchange sites

Developed mesotunnels

Thermo-alkali reaction

Inaccessible charge compensator ions

Si

Al

O

Ca

K

ACS Paragon Plus Environment

H