You are on page 1of 13

Journal of Hazardous Materials 244245 (2013) 444456

Contents lists available at SciVerse ScienceDirect

Journal of Hazardous Materials


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

Review

Adsorptive removal of hazardous materials using metal-organic frameworks (MOFs): A review


Nazmul Abedin Khan 1 , Zubair Hasan 1 , Sung Hwa Jhung
Department of Chemistry and Green-Nano Materials Research Center, Kyungpook National University, Daegu 702-701, Republic of Korea

h i g h l i g h t s
Metal-organic frameworks are very effective to remove hazardous materials. Mechanisms for adsorptive removal with MOFs were summarized. MOFs are surely regarded as potential adsorbents for clean environment.

g r a p h i c a l

a b s t r a c t

a r t i c l e

i n f o

a b s t r a c t
Efcient removal of hazardous materials from the environment has become an important issue from a biological and environmental standpoint. Adsorptive removal of toxic components from fuel, wastewater or air is one of the most attractive approaches for cleaning technologies. Recently, porous metalorganic framework (MOF) materials have been very promising in the adsorption/separation of various liquids and gases due to their unique characteristics. This review summarizes the recent literatures on the adsorptive removal of various hazardous compounds mainly from fuel, water, and air by virgin or modied MOF materials. Possible interactions between the adsorbates and active adsorption sites of the MOFs will be also discussed to understand the adsorption mechanism. Most of the observed results can be explained with the following mechanisms: (1) adsorption onto a coordinatively unsaturated site, (2) adsorption via acid-base interaction, (3) adsorption via -complex formation, (4) adsorption via hydrogen bonding, (5) adsorption via electrostatic interaction, and (6) adsorption based on the breathing properties of some MOFs and so on. 2012 Elsevier B.V. All rights reserved.

Article history: Received 3 September 2012 Received in revised form 29 October 2012 Accepted 4 November 2012 Available online 13 November 2012 Key words: Adsorption Adsorptive removal Hazardous materials Metal-organic frameworks Review

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. Common hazardous materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. Adsorptive removal of hazardous materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3. Metal-organic frameworks (MOFs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4. Purpose of this review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445 445 445 446 447

Corresponding author. Tel.: +82 53 950 5341; fax: +82 53 950 6330. E-mail address: sung@knu.ac.kr (S.H. Jhung). 1 Fax: +82 53 950 6330. 0304-3894/$ see front matter 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jhazmat.2012.11.011

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

445

2.

3.

Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Adsorptive removal of SCCs and NCCs from fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Adsorptive removal of organic contaminants from waste water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Adsorptive removal of heavy metal ions from efuent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Adsorptive removal of harmful gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

447 447 448 449 449 452 453 453

1. Introduction Presently, environmental pollution is one of the most problematic issues worldwide. There have been many trials both to reduce pollution and to eliminate the polluting materials from the environment. Common hazardous materials that exist in our environment are NOx , SOx , COx , H2 S, volatile organic compounds (VOCs), nitrogen-containing compounds (NCCs), sulfur-containing compounds (SCCs), dyes, pharmaceuticals and personal care products (PPCPs), and so on. On the other hand, porous metal-organic framework (MOF) materials are very interesting due to their versatile applications. MOFs are superior to other porous materials because of their high/tunable porosity, pore functionality, various pore architectures/compositions, open metal sites, and so on. Therefore, recently, extensive studies have been done on the adsorption/separation of various gaseous and liquid components with MOFs. In this review, the adsorptive removal of various toxic liquids and gases using virgin or modied MOFs will be discussed.

1.1. Common hazardous materials The most abundant hazardous components that exist in the environment can be classied into two categories based on their sources; these are naturally occurring hazardous materials and anthropogenic. A considerable amount of naturally occurring pollutants is present in the air, minerals, water, and soil. On the other hand, the anthropogenic pollutants generally originate from combustion, chemical reactions or from the unsecured efuent of toxic materials. The global energy demand is being supplied mainly by natural gas, coal, and crude oil. The burning of these energy sources emits huge amounts of toxic gases into the atmosphere. Generally, the emissions with the greatest concern for environmental air pollution are NOx , SOx , CO2 , VOCs, H2 S, NH3 , and other hydrocarbons [1,2]. Emissions of N2 O, SO2 , O3 , and CH4 enlarge the ozone levels in the troposphere and are also considered as greenhouse gases. Vehicle-related pollutants like SO2 , NO2 , CO, CH4 , and black carbon contribute to global warming [1]. The carbon balance of the world is one of the most important environmental issues currently; therefore, reducing anthropogenic CO2 emissions has become a great challenge for humanity. NH3 , one of the most widely used chemicals in laboratories and various industries including agricultural, manufacturing, refrigeration, etc., is responsible partly for the nitrogen containing pollutants in the atmosphere [35]. The American Conference Governmental Industrial Hygienists (ACGIH) has specied the allowable NH3 concentration as up to 25 ppm as the time-weighted average and 35 ppm as short-term exposure [5]. Recently, adsorption has been regarded as one of the most successful techniques for NH3 capture [58]. H2 S, another toxic and odorous air pollutant has also been removed through sorption [911]. VOCs are chemicals with high vapor pressure, generally emitted from solvents, resins, paints, adhesives, etc. [12,13]. The common hazardous VOCs are benzene, naphthalene, toluene, phenolics, xylenes, and so on. VOCs are considered hazardous materials since they are harmful to the environment and human health.

Organic compounds containing sulfur or nitrogen are naturally occurring species that are present in fossil fuels and oils like crude oil, gasoline, diesel, jet fuel, and heating oil. More than 85% of the global energy has originated from fossil fuels [14]. The combustion of a huge amount of fossil fuel is the major source of toxic emissions that account for the dangerous air pollution, greenhouse effect, and global warming as well as the harmful impact on living organisms [2]. Therefore, recently, it has become a great challenge to remove these harmful compounds from crude oils before utilization. For example, based on EU and US guidelines, the sulfur level in fuels should be less than 10 ppmw and 15 ppmw, respectively [1,1517]. The most abundant pollutant of water is dye materials. Dyes are widely used in the textile, leather, paper, painting, and plastic industries. Around 100,000 commercially available dyes are produced at a rate of 7 105 tons per year [18] with 2% of the produced dyes being discharged into aquatic systems as efuent [18]. Removal of these materials from waste water is very important because water quality is highly inuenced by color [18] and even a small amount of dye is highly visible and considered to be toxic and extremely hazardous to aquatic living organisms [1820]. PPCPs include a class of chemical contaminants that exist in water even after these products are utilized for medicines, cosmetics, fragrances, veterinary drugs, fungicides, disinfectants, and agricultural practices [2123]. The presence of these kinds of PPCPs in the efuent of wastewater treatment plants, rivers, lakes and occasionally, in groundwater, has been demonstrated by the recent research [24]. Moreover, it is reported that, a class of PPCPs is unsafe for living organisms and may cause endocrine disruptions changing hormonal actions [25,26]. The disposal of heavy metal ions in processed water is still a considerable amount. Some metals like lead (Pb), arsenic (As), copper (Cu), mercury (Hg), antimony (Sb), chromium (Cr), manganese (Mn), and cadmium (Cd) are signicantly toxic to ecological systems and human beings [2731]. The recovery of those harmful metal ions from environment has been a global concern for the last few decades [2731]. 1.2. Adsorptive removal of hazardous materials Adsorption has been considered to be superior to other techniques for decontamination in view of its comparatively low cost, wide range of applications, simplicity of design, easy operation, low harmful secondary products and facile regeneration of the adsorbents. Adsorptive removal is based on the ability of a porous adsorbent to selectively adsorb some specic compounds from the atmosphere or renery streams. The compounds, which have a suitable size and shape, can be removed via adsorption since these compounds have easy access to the pores of the solid sorbents. Based on the types of interactions between an adsorbate and a porous sorbent, the adsorption can be categorized as a physical or chemical one [1,32]. Physical adsorption is usually called adsorptive adsorption, whereas chemical adsorption is called reactive adsorption. In the case of adsorptive removal, the adsorbates are generally trapped with weak (van der Waals) forces inside the pores of the solid adsorbents. Therefore, the adsorbent can be easily

446

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

Fig. 1. Widely used porous sorbents: (a) activated carbon, (b) HY (FAU) zeolite, (c) MCM-41 and (d) MIL-101 (Cr).

regenerated by simple solvent exchange or some other physical treatments like calcination and sonication. On the other hand, reactive adsorption occurs through the formation of true chemical bonds between the adsorbates and the adsorbent. Regeneration of the spent adsorbent is usually carried out with chemical treatments. The efciency of the adsorptive removal is determined mainly by the adsorption capacity of the adsorbents, selectivity for specic compounds, durability, and regenerability of the adsorbents. Various porous adsorbents (Fig. 1) such as activated carbons [3337], zeolites [18,3840], mesoporous materials [4144], and MOFs [14,15,4548] have been studied for adsorptive removal of hazardous compounds. For efcient adsorptive removal, porosity, pore geometry [46,49], and specic adsorption sites are required [15,50,51]. Additionally, some active species like various functional groups (acidic or basic), metal ions, metal oxides, metal salts, and polyoxometalates are usually incorporated into the porous adsorbents for selective adsorption of hazardous components through common interactions like acid-base, -complexation, interaction, and hydrogen bonding. Among the modication techniques, post synthetic modication, functionalization, ion exchange, impregnation, and loading of porous adsorbents have been widely studied. 1.3. Metal-organic frameworks (MOFs) MOFs are basically composed of two major components: a metal ion or a cluster of metal ions and an organic molecule called a linker. The organic units are typically di-, tri-, or tetradendate ligands [52,53]. Some typical MOFs are shown in Fig. 2. In the past, porous hybrid frameworks were called coordination polymers [53] since the inorganic part contained either an isolated polyhedral or small clusters like in coordination chemistry. However, very soon it was found that porous hybrid solids could possess inorganic parts with a larger dimensionality which could give rise to chains (1D), layers (2D), and even frameworks (3D) that were called MOFs [54].

Moreover, MOFs were further extended such as IRMOFs (Isoreticular MOFs), MMOFs (microporous MOFs), PCPs (porous coordination polymers), and so on. MOF materials have many advantages compared with zeolitetype materials. Zeolite-related inorganic hybrid materials need an organic or inorganic template to form; however, in the case of MOFs, a solvent is the main templating molecule [53,55]. Another important feature is that most of the metal cations can participate in MOF formation [53] compared with inorganic materials which are based on a few cations such as Si, Al, and P. MOFs have signicant advances for the formation of a series called isoreticular MOFs, by only changing the length of the ligands with the same metal species in which the pore size of the corresponding frameworks depends on the length of the ligands [56,57]. Moreover, several analogous MOFs can also be prepared from identical ligands and different metallic components [57]. The interest in MOF-type materials is due to the huge porosity and easy tunability of their pore size and shape from microporous to mesoporous scale by changing the connectivity of the inorganic moiety and the nature of the organic linkers [58,59]. Moreover, MOFs have many potential applications including adsorption/storage of carbon dioxide [14,48,60], hydrogen storage [61], adsorption of vapours [62], separation of chemicals [63], drug delivery/biomedicine [64], polymerization [65], magnetism [66], catalysis [67], luminescence [68], and so on. Not only researches on MOFs but also the number of publications (per year) related to MOFs has been increased rapidly [45] because of potential applications of MOFs in various elds. MOFs are promising materials for adsorption-related applications because of the easy modication of their pore surfaces which leads to the selective adsorption of some specic guest molecules having particular functional groups. Recently, MOF-type materials have also been investigated widely for the adsorptive removal of various hazardous materials from the environment due to their huge porosity and pore geometry [6971]. Moreover, the central metals [7275], coordinatively unsaturated sites (CUS or open metal sites) [7,74,7678],

Fig. 2. Structures of typical metal-organic frameworks. (a) MOF-5, (b) Cu-BTC and (c) CPO-27. Source: Reproduced with permission from ref. [60]. Copyright 2012 American Chemical Society.

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

447

functionalized linkers [7,30,79,80], and loaded active species [4,81,82] have been employed successfully for some additional interactions between the adsorbates and MOF materials; that make MOFs superior to other porous adsorbents for efcient adsorptive removal of hazardous compounds. Among the various host-guest interactions, acid-base [72,82,83], -complexation [81,84], H-bonding [7,85] and coordination with open metal sites [7,74,7678] play important roles in the preferred adsorption. 1.4. Purpose of this review This review will mainly focus on the adsorptive removal of common hazardous materials that exist in the environment as gas or liquid with widely studied porous MOFs. The advantageous characteristics of MOFs that play a key role in the adsorption/removal of particular hazardous compounds will be pointed out. The effects of central metal ions, open metal sites, linkers, porosity, functionalization/modication of MOFs in adsorption, and the possible interactions between adsorbate and adsorbent will be discussed. Moreover, adsorption parameters like adsorption kinetics and thermodynamics will also be included to understand the adsorption phenomena. 2. Discussion 2.1. Adsorptive removal of SCCs and NCCs from fuel SCCs and NCCs are widely known contaminants in petroleum rening and in fuels. Removal of these compounds before commercial use is extremely important because of environmental protection and catalyst poisoning issues. The removal of SCCs and NCCs from liquid fuels by adsorption using porous MOFs has been explored to date [1,15,17,72,8184,8691]. Cychosz et al. [90] studied a number of MOFs having different pore sizes and shapes for the adsorption of various SCCs in model oils and found that MOFs are superior to other porous adsorbents like zeolites or activated carbon due to the selectivity towards SCCs (Fig. 3). Peralta et al. has shown similarly that MOFs with CUSs (such as copper trimesate or Cu-BTC [92] and nickel 2,5-dihydroxyterephthalate or CPO-27(Ni) [93]), contrary to zeolites, exhibit stronger afnity for thiophene than that for toluene [94]. Achmann et al. [17] also reported successful removal of thiophene and tetrahydrothiophene from model oils by using Cu-BTC. Blanco-Brieva et al. [78] showed that CuBTC adsorbs a considerable amount of DBT at ambient temperature (304 K) which was eight times higher than that of bench-marked Ytype zeolite or activated carbons. The higher uptake was explained in terms of a stronger interaction of the S-atom of DBT with the surface Cu2+ ions of the Cu-BTC framework. Park et al. [95] synthesized isostructural MOFs having N-substituted linkers, and showed that the uptake of SCCs was increased with the introduction of electron-decient linkers. The authors explained the high uptake with an increased electronic interaction between electronrich SCCs and electron-decient linkers [95]. Selective adsorption of DBT from solutions containing iso-octane, naphthalene and benzene by Mo(CO)6 decomposed onto the surface of zinc terephthalate or MOF-5 (Mo loadings up to 20 wt%) has been reported by Shi et al. [87]. The breakthrough approached 0.5 mmol S g1 suggesting a strong afnity between DBT and the Mo carbides and/or oxycarbides of the adsorbent. The high concentrations of aromatics like naphthalene and benzene caused the decrease in the sulfur adsorption capacities. Khan et al. [72] studied adsorption kinetics and thermodynamics in the adsorption of benzothiophene (BT) over three analogous MOFs such as metal terephthalates (MIL-53s(Al, Cr) and MIL-47(V); MIL stands for the Material of Institut Lavoisier) [9698]. They

Fig. 3. Adsorption isotherms of SCCs over various MOFs: benzothiophene (top), dibenzothiophene (middle) and 4,6-dimethyldibenzothiophene (bottom). Source: Reproduced with permission from ref. [90]. Copyright 2008 American Chemical Society.

concluded that the central metal ions of MOFs play an important role in the adsorption of BT from liquid fuels. MIL-47 showed the highest adsorption capacity and the fastest adsorption kinetics among the analogous MIL-53s because of the high acidity (Fig. 4). Therefore, an acid-base interaction (between the acidic MIL-47 and slightly basic BT) was suggested for the favorable adsorption. The driving force of BT adsorptions over the adsorbents in that study was due to an increased entropy change rather than an enthalpy change. Moreover, introduction of an acidic component to porous materials also results in a specic adsorption of slightly basic SCCs through acid-base interactions [82,83]. Phosphotungstic acid (PWA)-loaded porous Cu-BTC was examined by Khan and Jhung [82] to understand the effect of PWAs on the adsorption/removal of BT. The maximum adsorption capacity (Q0 ) increased with increasing PWA loading up to a W/Cu (wt/wt) ratio of 0.22 in PWA/Cu-BTCs, resulting in an increase in the Q0 by 26% compared with the virgin Cu-BTC. Since there was no remarkable change in the surface area and pore volume for the virgin and PWA loaded Cu-BTCs, it was suggested that the improved Q0 over PWA/Cu-BTCs was due to favorable interactions like acidbase ones between the acidic

448

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

Fig. 4. Adsorption isotherms for benzothiophene adsorption over the MIL-47(V) and MIL-53s at 298 K [72].

PWA and slightly basic BT. Therefore, the acidity or basicity of MOFs originates either from the central metals or from incorporated species might be involved in adsorbing compounds through acid-base interactions. It has been reported that metal ions like Cu+ , Ag+ , Pd2+ and 2+ Pt have an adsorption capability for SCCs through -complex formation [99101]. Yang and coworkers have developed complex-based adsorbents for desulfurization [99]. In most cases, -complexation adsorbents have been prepared by high temperature calcination of metal-ion-exchanged zeolitic materials [99]. Additionally, Cu2 O-loaded porous materials like Al2 O3 , MCM-41, and SBA-15 have also been reported for the adsorptive removal of SCCs through -complex formation between Cu+ and SCCs [102,103]. Khan and Jhung [81] reported a remarkable adsorptive performance for BT adsorption with CuCl2 -loaded-MIL-47 which involved reducing Cu2+ to Cu+ (by the V3+ of the MIL-47) at ambient condition for an efcient -complex formation with SCCs in n-octane (Fig. 5). The authors demonstrated that the high adsorption capacity of CuCl2 /MIL-47 (310 mg BT g1 ; 122% of CuI Y [104] that had the highest capacity so far) was mainly due to a synergy between the reduced Cu+ ions (for -complexation) and MIL-47. On the other hand, CuCl2 -loaded-MIL-53s(Al and Cr) did not show any benecial effect of CuCl2 loading on BT adsorption [105]. Different to V3+ , Al3+ and Cr3+ were not efcient in the selective reduction of Cu2+ to Cu+ ; therefore, it can be assumed that the central

metals of MOFs as well as their oxidation states are also important in the various potential applications of MOFs. Moreover, Cu+ loaded iron terephthalate, MIL-100(Fe), was also evaluated by the same group [84] as an efcient adsorbent for BT adsorption to investigate the benecial effect of -complex formation. The Cu+ species was incorporated facilely onto the porous MIL-100(Fe) under mild condition through a one-step synthesis of Cu2 O (without selective reduction of Cu2+ to Cu+ at about 923 K). Q0 increased with increasing copper loading up to a Cu/Fe (wt/wt) ratio of 0.07 in Cu+ -loaded-MIL-100(Fe), resulting in an increase in the Q0 by 16% compared with the virgin MIL-100(Fe). Compared to SCCs, very few reports have been published for the adsorptive removal of NCCs with MOFs. In 2010, Nuzhdin et al. [89] used highly porous chromium terephthalate, MIL-101(Cr) to remove various NCCs from light cycle oil, and a maximum adsorption capacity of 19.6 mg N g1 was reached. They attributed the high nitrogen adsorption capacity to the coordination of the NCCs on the Cr3+ sites (CUS) of the MIL-101(Cr). They observed that the better the steric accessibility of the N atom of the substrate molecule was, the higher the adsorptive performance of the MOF was. A more detailed report was published by Maes et al. [106] in which they examined a series of MOFs for the adsorption of a mixture of NCCs and SCCs from liquid fuel. Selective adsorption was observed with several MOFs such as MIL-100s(Fe, Cr, Al) and MIL-101(Cr) and the adsorption phenomena were nicely explained with Pearsons HSAB (hard/soft acid/base) concept. According to this concept, harder bases, such as NCCs, interact preferentially with hard Lewis acid sites such as Fe3+ , Cr3+ , and Al3+ and also with intermediate Lewis acid sites. On the other hand, softer bases like SCCs have a preference to interact with intermediate or soft Lewis acid sites such as Cu2+ , Zn2+ , Co2+ , Ni2+ , and Cu+ . Ahmed et al. [83] studied the detailed acid-base effect on adsorption of NCCs over virgin and modied MIL-100(Cr). The MOF, MIL-100(Cr), was modied by grafting compounds having acidity or basicity onto the CUS of the MIL-100(Cr). It was observed that the adsorptive removal of basic quinoline was improved with an acid-grafted MOF; however, adsorption was severely decreased by a basic-group-grafted MOF because of acid-base interactions.

2.2. Adsorptive removal of organic contaminants from waste water Several virgin or modied MOFs have been tested for the removal of various organic contaminants from waste water. Generally, pore structure [107], charge interaction between the

Fig. 5. Adsorption mechanism (left) and adsorption isotherms (right) of benzothiophene over virgin and CuCl2 -loaded MIL-47(V) [81].

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

449

adsorbent and adsorbate [23,108,109], open metal sites [77] or the breathing properties [110] of the MOFs have been reported as important parameters or mechanisms for adsorption. Haque and coworkers [108] reported that MOF-235 (another phase of iron terephthalate [111]) could be used for the removal of harmful dyes (anionic dye methyl orange (MO) and cationic dye methylene blue (MB)) through adsorption. The adsorption capacities of MO and MB with MOF-235 were 477 and 187 mg g1 , respectively, and only 11 and 26 mg g1 for MO and MB, respectively, with activated carbon. The adsorption rates for the adsorptions were also much faster with MOF-235 than with the activated carbon. They also conrmed that, the adsorption of MO and MB dyes was highly dependent on the pH of the solution similar to other reports [18,77,80,109]. The density of the positive charge of the adsorbent decreases with increasing pH of the MO solution; therefore, decreased adsorption was observed with increasing pH. On the contrary, the concentration of the negative charge of the adsorbent increases with increasing pH resulting in increased MB adsorption. Adsorption of MO and MB were spontaneous and endothermic, and the entropy (the driving force of the adsorption) increases with the adsorption of MO and MB. The authors suggested that the endothermic adsorptions occurred because of a stronger interaction between the pre-adsorbed water and the adsorbent than the interaction between the MO or MB and the adsorbent. Simple electrostatic interaction was also shown between PPCPs and MOFs by Hasan et al. [23] for the liquid phase adsorption of naproxen and clobric acid. The adsorption capacities of two different MOFs, MIL-101(Cr) and MIL-100(Fe) were compared with conventional adsorbents such as activated carbon. The efciency decreased in the order of MIL-101(Cr) > MIL-100(Fe) > activated carbon both for the adsorption rate and adsorption capacity. Large surface area or pore volume was benecial for high adsorption, and low solution pH was proven to be favorable for the adsorption of PPCPs over MIL-101(Cr). The open metal sites or coordinatively unsaturated sites (CUS) of MIL-100(Fe) were the active species to bind malachite green (MG) [77]; whereas, interaction between the Lewis base N(CH3 )2 in MG and the CUS (Lewis acid) of MIL-100(Fe) occurred due to the replacement of the water molecules by the Lewis base N(CH3 )2 of MG. The possibility of a interaction between the benzene rings in MG and MIL-100(Fe) over the adsorption was also suggested. Moreover, the adsorption of MG on MIL-100(Fe) was an endothermic process and the driving force for the adsorption was controlled by an entropy effect rather than an enthalpy change. Huang et al. [107] synthesized hierarchically mesostructured MIL-101(Cr) by using cetyltrimethylammonium bromide (CTAB) as a surfactant, and the material showed remarkably accelerated adsorption kinetics for liquid phase adsorptive removal of MB. They observed that all MB molecules with 30 ppm of initial concentration were adsorbed onto the hierarchically mesostructured MIL-101(Cr) in about 110 min. Maes et al. [110] reported the liquid phase adsorption of phenol from contaminated water over MIL-53(Cr) which has a characteristic breathing property of the framework. Because of this breathing effect, a stepwise increase in the adsorbed amount with increasing concentrations of phenol was observed. This abnormal observation was explained with a sudden expansion of the pores, and nally allowed more phenol to be adsorbed. In the case of p-cresol adsorption, this effect was not observed which might be due to the difference in polarity between phenol and p-cresol. MOFs having a characteristic breathing property also showed a noticeable effect in the vapor phase adsorption of alkanes [112]. Functionalization of MOFs has also been reported for the adsorptive removal of charged liquid contaminants from waste water through electrostatic interactions. MIL-101(Cr) after being modied to have a positive charge has efciently been used to adsorb anionic MO through charge interaction [80]. Fig. 6 shows the

interaction sites of anionic MO and functionalized MIL-101(Cr) adsorbents. The adsorption capacities and kinetic constants were in the order of activated carbon < MIL-101(Cr) < ethylenediamineMIL-101(Cr) (or ED-MIL-101(Cr)) < protonated grafted ethylenediamine-grafted MIL-101(Cr) (or PED-MIL-101(Cr)), and the adsorption capacity of PED-MIL-101(Cr) was 194 mg g1 [80]. The kinetic constant over PED-MIL-101(Cr) was around 10 times larger than that of the activated carbon under the experimental conditions. The number of desorbed water molecules was larger than that of the adsorbed MO molecules (MO is very bulky compared with water; therefore, several water molecules may be desorbed by the adsorption of one MO molecule); consequently, the randomness of the adsorption of MO increased. Therefore, the high adsorption of PED-MIL-101(Cr) was explained by an entropy effect (large positive S) [80]. 2.3. Adsorptive removal of heavy metal ions from efuent Numerous techniques are available for metal recovery from waste water. Many of these are established methods [113118], while others are still in the experimental stages. Recently, MOFs have been regarded as robust adsorbents for the adsorptive removal of heavy metal ions from efuent [30,31,119]. Zhu et al. [31] reported the removal of arsenic (As5+ ) from aqueous solutions over MIL-100(Fe). According to that report, the adsorption capacity of As5+ with MIL-100(Fe) was around 6 and 36 times higher than that of iron oxide nanoparticles and commercial iron oxide powders, respectively (Fig. 7). MIL-100(Fe) was able to absorb As5+ over a wide range of pH (212). However, at a pH above 12, the adsorption capacity decreased drastically since MIL-100(Fe) is not stable at high pH. To shed light on the adsorption mechanism they conducted XPS, IR and TEM analysis and conrmed that the adsorption took place via formation of Fe-O-As bonds and As5+ was preferentially adsorbed onto the interior of the MIL-100(Fe) rather than the outer surface. As a result, porous MIL-100(Fe) provides more interior spaces compared to Fe2 O3 nanoparticles, which resulted in a six fold higher adsorption capacity. Ke et al. [30] reported Hg2+ adsorption from water over thiol functionalized Cu-BTC. Functionalization was done through coordination bonding between the thiol groups of dithioglycol and the CUS of Cu-BTC. Thiol functionalized Cu-BTC showed a very high adsorption capacity (714 mg g1 ) for Hg2+ whereas the virgin Cu-BTC showed no adsorption for Hg2+ under the same experimental condition. The authors revealed that the very high adsorption of Hg2+ was due to the high density thiol groups on the inner surface of the porous MOFs having a huge specic surface area and a high density of adsorption sites. Moreover, it has been suggested that MOFs can be utilized for the removal of metal ions with the ion-exchange of MOFs [120,121]. Additionally, in some cases, simple ion-exchange procedures lead to not only metal ions removal but also formation of a few isostructural MOFs [122,123]. 2.4. Adsorptive removal of harmful gases Harmful or toxic gases/vapors such as H2 S, SO2 , NH3 , CO2 , CO, NO, benzene have been adsorbed over various MOFs [57,48,60,63,6971,7476,85,124152]. Most of these gases/vapors are released as waste byproducts from various industries into the environment. Effective capture of these unsafe gases/vapors is very important for a clean environment. So far, for efcient adsorptive removal, not only the pore size/porosity but also the 2 -OH, 2 -O, open metal sites, and functional groups of adsorbents are all quite promising for some specic interactions between hazardous materials and the host adsorbents. Hamon et al. [75] did a comparative study of MOFs that are stable and easily regenerable upon H2 S sorption through pressure swing

450

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

Fig. 6. Proposed electrostatic interaction between methyl orange and adsorbents [80].

adsorption processes. The polar H2 S molecules interact strongly with 2 -OH of the inorganic chain of MIL-53s(Al, Cr), leading to a closure of the pores at low pressure. Interestingly, with increasing pressure, the pores were reopened with the breaking of the strong H2 S. . .HO (on the metal site of MOFs) interactions, and consequently, all the pores were lled through weak host-guest interactions which resulted in steps in the adsorption isotherm. The rst and second steps began at 4.5, 118 kPa and 9.0, 210 kPa for MIL-53(Al) and MIL-53(Cr), respectively. The adsorption isotherms of H2 S adsorption over the investigated MOFs are shown in Fig. 8. The maximum adsorption capacities of H2 S sorption were reached to 13.12 and 11.77 mmol g1 for MIL-53(Al) and MIL-53(Cr), respectively, at high pressure (1.6 MPa). On the other hand, MIL-47,

analogous to MIL-53s but with no OH group, exhibits a type-I isotherm upon H2 S sorption. The adsorption capacities of MIL-47 and MIL-53s(Al, Cr) at high pressure were similar; therefore, it was suggested that the pores in MIL-53s reopened at high pressure. Moreover, the authors also reported a huge uptake of H2 S with the large-pores of MIL-100(Cr) and MIL-101(Cr) exhibiting type-I shaped adsorption isotherms; however, the adsorptions appeared to be irreversible. The maximum adsorbed quantities at 2 MPa were 16.7 and 38.4 mmol g1 for MIL-100(Cr) and MIL-101(Cr), respectively. Partial destruction of the framework or strong interactions between the framework and the H2 S molecules were suggested for the irreversible adsorption phenomenon. Moreover, it was also reported [85] that the adsorption of H2 S preferentially occurs

Fig. 7. (a) Adsorption isotherms and (b) linearized Langmuir isotherms for As5+ adsorption by MIL-100 (Fe) gel, Fe2 O3 nanoparticles and bulk Fe2 O3 powders (pH 4, m/V = 5.0 g/L, T = 298 K). Source: Reproduced with permission from ref. [31]. Copyright 2012 American Chemical Society.

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

451

Fig. 8. Excess adsorbed quantities of hydrogen sulde on MIL-53(Cr) (pink), MIL53(Al) (blue), and MIL-53(Fe) (yellow) at 303.1 K at pressures up to 2.0 MPa. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.) Source: Reproduced with permission from ref. [75]. Copyright 2009 American Chemical Society.

through the formation of the hydrogen bond between the 2 -O atom of the V O V moiety in MIL-47 and the H2 S molecules, where H2 S acts as a hydrogen donor. Interaction of acidic H2 S with other basic centers in MIL-47 such as the oxygens of the carboxylate group and electrons of the benzene ring was also suggested [85]. A uorinated metal-organic framework (FMOF-2, obtained from 2,2 -bis(4-carboxyphenyl)hexauoropropane and zinc nitrate hexahydrate) having breathing characteristics was also reported for the adsorptive removal of toxic acidic gases [124]. FMOF-2 was quite stable for the adsorption of SO2 and H2 S. At room temperature and 1 bar, the calculated weight capacities for SO2 and H2 S were 14.0% and 8.3%, respectively, with FMOF-2. The open metal sites of MOFs have been reported several times as the active sites for the adsorptive removal of various toxic gases. Petit and Bandosz [125] reported composites of MOFs (MOF-5, Cu-BTC or MIL-100(Fe)) and a graphitic compound (graphite or graphite oxide, GO) for the adsorptive removal of NH3 , H2 S and NO2 under ambient conditions. The open metal sites of porous MOFs coordinated with the oxygen groups of GO led to the formation of a new pore space in the interface between the carbon layers and the MOF units to form composites with distinct properties. More than 12% (for NH3 ), 50% (for H2 S) and 4% (for NO2 ) increases in the adsorption capacity were observed using a GO/Cu-BTC composite (compared with the virgin Cu-BTC). The enhancement of the adsorption capacities for the toxic gases was explained with the

new porosity created in the interface where the dispersive forces were the strongest. Additionally, reactive adsorption was also suggested since the lone pair electron of NH3 , H2 S and NO2 coordinates to the CUS of the MOFs and breaks the framework structure, and nally results in metal salts [125]. Britt et al. [7] demonstrated the potential applicability of six MOFs, composed of a Zn4 O(CO2 )6 cluster linked by terephthalate (MOF-5), 2-amino terephthalate (IRMOF-3), benzene-1,3,5-tris(4benzoate) (MOF-177), diacetylene-1,4-bis(4-benzoic acid) (IRMOF62), Zn2 O2 (CO2 )2 chains linked by 2,5-dihydroxyterephthalate (CPO-27) and Cu-BTC in the adsorption and separation of several harmful gases or vapors including SO2 , chlorine, NH3 , benzene, ethylene oxide, tetrahydrothiophene, and the results were compared with BPL carbon. Fig. 9 shows the breakthrough curves of SO2 and NH3 adsorption on various MOFs. Several factors such as the open metal sites of CPO-27 (or M2 (dobdc)(H2 O)2 ; H4 dobdc = 2,5dihydroxyterephthalic acid) or Cu-BTC and the active adsorption site with particular functional groups (such as NH2 in IRMOF-3) were proven to play an important role in determining the dynamic adsorption performance of these MOFs. For SO2 adsorption, CPO-27 showed the best performance over the other MOFs and its capacity was 6 times larger than that of the BPL carbon. This favorable adsorption is due to the presence of a highly reactive 5-coordinate zinc species along with the potentially reactive oxo group in the CPO-27. On the other hand, Cu-BTC showed a high efciency equal to or greater than that of BPL carbon for all the gases tested except for Cl2 since it does not naturally act as a ligand. In the case of NH3 , the presence of the NH2 in IRMOF-3 sharply improves the adsorptive performance compared to the virgin IRMOF-1 or MOF-5. Before breakthrough, IRMOF-3 adsorbed almost 71 times as much NH3 compared with BPL carbon, and this can be explained with the capability of NH3 to readily form hydrogen bonds. Glover et al. [74] also showed adsorptive removal of several harmful gases including NH3 , CNCl, SO2 , and octane vapor using M-CPO-27 (M Zn, Co, Ni, or Mg) in both dry and humid conditions. Here, the experimental breakthrough results revealed that all these MOFs, with open metal sites, are capable of adsorbing the studied toxic gases only in dry conditions, while in humid conditions the adsorption capability was reduced dramatically since the competitive adsorption of water started to dominate. The decrease in adsorption capacity was not very noticeable in the case of NH3 gas, i.e. ammonia was adsorbed both in dry and humid conditions. Generally, it can be concluded that water vapor has a negative effect on the adsorption of various gases over CPO-27 type materials. Dathe et al. [129] demonstrated adsorption of SO2 over Ba(CH3 COO)2 (or Ba(NO3 )2 and BaCl2 )-impregnated Cu-BTC. Impregnation led to small microcrystals of barium salts in the pores of Cu-BTC while partial destruction of host structure was observed in the case of only BaCl2 . The authors pointed out that at high temperature, SO2 uptake exceeded the stoichiometric capacity based

Fig. 9. Selected kinetic breakthrough curves of (A) SO2 and (B) NH3 gases over various MOFs [7].

452

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

on the Ba2+ concentration. As a result, the excess SO2 uptake is due to the contribution of chemical bonding between the metal cations (from the MOF) and SO2 , which ultimately resulted in the formation of Cu-sulfates. Therefore, at low temperature, Cu-BTC behaved as a good host material to possess highly dispersed barium salts. On the contrary, at high temperatures, Cu-BTC decomposed and produced isolated Cu species that acted as SOx storage sites forming nally Cu-sulfates. Irreversible SOx storage, however, is the main disadvantage of barium salts-impregnated Cu-BTC materials. Cu-BTC has been widely studied for the selective adsorption of CO2 molecules [14,60,132,133,136]. Yazaydn et al. [135] showed striking enhancements in the abilities to capture CO2 as well as enhanced CO2 selectivity over N2 and CH4 with a Cu-BTC framework containing 4 wt% water molecules coordinated to the open metal sites of the framework. It was suggested that the Coulombic interactions between water and CO2 are mostly responsible for the enhancement in CO2 adsorption while the quadrupole moment of CO2 interacts with the electric eld gradient of the sorbent, which is increased when water occupies the copper open metal site. It may be one positive example of water; however, the effects of water vapor on various adsorptive performances with other MOFs having open metal sites require further studies because water usually hinders adsorptions. The effect of the metal center on the adsorption capacity and selectivity of CO2 adsorption was studied by Dietzel et al. [136] with a series of isostructural MOFs M-CPO-27s (M Ni, Co, Zn, Mg, Mn). It was reported that at 298 K and high pressure (50 bar), the CO2 uptakes were 51 wt% for CPO-27(Ni) and 63 wt% for CPO-27(Mg). Caskey et al. [73] showed 30.6 and 35.2 wt% CO2 uptakes at 1 atm with CPO-27(Co) and CPO-27(Mg), respectively. The CO2 uptake of CPO-27(Mg) is much higher than that of any other member of the series [73,136,137]. CPO-27(Mg) takes up to 8.9 wt% CO2 before breakthrough, corresponding to 0.44 molecules of CO2 per Mg ion at low pressure [137]. Thus, it is quite optimistic that CPO-27(Mg) represents a major advance in the CO2 separation capacity of MOFs. The end-on coordination mode for CO2 , with the increased ionic character of the Mg2+ O interaction, was subjected to the high adsorption capacity of CPO-27(Mg) [73,137,138]. Ahn et al. reported the improved CO2 adsorption with high quality and small sized adsorbents obtained by sonication or microwave heating [139141]. Small and homogeneous particles of CPO-27(Mg) synthesized via sonication adsorbed 350 mg g1 of CO2 at 298 K and showed high isosteric heat of adsorption (4222 kJ mol1 ) [141].The MOFs having amine functionalized linkers showed enhanced afnity towards CO2 [60,70,79,132,142146]. It is obvious to expect enhanced CO2 adsorption with amine functionalized MOFs through the acid-base or electrostatic interactions. Moreover, strong interaction between the amino functionality and CO2 has also been reported [143,144]. A remarkable CO2 uptake with MOFs that have an ultrahigh surface area was reported by Furukawa et al. [69]. MOF-210 [Zn4 O(BTE)4/3 (BPDC) in which BTE: 4,4 ,4 -(benzene-1,3,5-triyl-tris(ethyne-2,1diyl))tribenzoate and BPDC: biphenyl-4,4 -terephthalate] and MOF-200 [Zn4 O(BBC)2 (H2 O)3 in which BBC: 4,4 ,4 -(benzene1,3,5-triyl-tris(benzene-4,1-diyl))tribenzoate] having BET surface areas of 6240 and 4530 m2 g1 , respectively, were able to uptake 2400 mg g1 of CO2 at 50 bar which exceeds the uptake values of other highly porous MOFs like MOF-177 and MIL-101(Cr)c, and so on [70,134]. It was assumed that unlike hydrogen and methane uptakes, the uptake capacities for excess CO2 are directly related to the total pore volume of the highly porous MOFs. CO is another toxic gas that is present in the environment. Because of its strong binding ability with metal sites, MOFs having CUS could adsorb CO selectively over other gases. A molecular simulation study revealed that Cu-BTC was quite a selective adsorbent for CO over H2 and N2 at 298 K [76]. It was suggested that

Fig. 10. Adsorption isotherms of benzene over various adsorbents at 303 K [144].

the electrostatic interaction between the partial charge of CUS of Cu-BTC (Cu2+ sites) and the CO dipole dominated the adsorptive performance. NO is another gas molecule which has great interest to remove/separate for environmental applications. A number of MOF materials have been employed successfully to adsorb/separate the gas via adsorption [131,148151]. Xiao et al. reported the gravimetric adsorption of NO on the open metal sites of Cu-BTC at 196 K (1 bar) and found that around 9 mmol of NO was adsorbed over 1 g of Cu-BTC which was signicantly higher than any other reported porous solids for the adsorption of NO [148]. Both vapor phase and liquid phase adsorption of benzene over chromium terephthalate, MIL-101(Cr), was studied by Jhung et al. [152]. MIL-101(Cr) adsorbed 16.7 mmol g1 of benzene in vapor phase, and this value was almost 5.5, 8.7 and two times larger than SBA-15, H-ZSM-5 and a commercial activated carbon, respectively (Fig. 10). The adsorption capacity of MIL-101(Cr) for benzene exceeded the adsorption capacity of pitch-based activated carbon (12.4 mmol g1 ) with a large surface area of 26003600 m2 g1 , and this high adsorption capacity was explained by the high porosity of MIL-101(Cr). Moreover, in the case of liquid phase adsorption in water, MIL-101(Cr) synthesized by microwave irradiation [153155] showed better performance compared to a commercial activated carbon and one MIL-101(Cr) synthesized by conventional hydrothermal techniques [152]. The potential applications of chromium terephthalate, MIL-101(Cr), in vapor phase adsorption of VOCs such as ethyl acetate and p-xylene were also conrmed by Li and coworkers [71,156,157]. 3. Summary Adsorption plays a signicant role in removing hazardous materials both from liquid and gas phases; therefore, it has attracted considerable attention in both scientic research and commercial applications. The invention of new materials for adsorption applications is indispensable interest and surely will be continued further. As a class of recently developed porous materials, MOFs have already shown huge potential applications in gas/vapor and liquid phase adsorption/removal of hazardous materials such as SCCs, NCCs, dyes, PPCPs, phenolics, SOx , NOx , VOCs and so on. MOFs are superior to other porous sorbents in adsorptive removal of various toxic components because of their high surface area, various pore geometries, facile functionalization, and tunable porosities. The variety of central metals in the framework of MOFs has led to a new strategy towards adsorption of various hazardous compounds like SCCs, NCCs, CO2 , SO2 and so

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456

453

Scheme 1. Important mechanisms for the adsorption of hazardous materials over MOFs.

on through specic interactions. For example, the coordinatively unsaturated metal sites (or open metal sites) of MOFs have been used successfully to bind (or coordinate) various polarizable toxic compounds like SCCs, NCCs, CO2 , CO, H2 S, NH3 and SO2 . Acid-base properties of MOFs showed a dominant role in the adsorption of hazardous contaminants having opposite acidic-basic properties. Moreover, incorporation of active acidic/basic species and metals into MOFs dramatically improves the adsorption capacity for various toxic compounds through acid-base interaction and -complex formation, respectively. Sometimes simple electrostatic interactions between MOF materials and its counter anionic or cationic adsorbates are also believed to play a dominating role in adsorption of anionic or cationic dyes, PPCPs over different MOFs or modied MOF materials. The hydrogen bonding between 2 -OH or 2 -O of MOFs inorganic chain and guest molecules has been suggested for the efcient adsorptive removal of hazardous materials. The unusual breathing behavior of porous MOFs such as MIL-53s is quite interesting for pressure swing adsorption of polar molecules like CO2 , H2 S and SO2 . Important mechanisms for the adsorption of hazardous materials over MOFs are summarized in Scheme 1. This review reveals that MOFs are undoubtedly regarded as potential adsorbents in the eld of liquid and gas phase adsorptions/removals. Moreover, the versatility in the characteristics of MOFs materials can be imparted through simple modications or functionalizations that allow for a widespread applicability of these materials. Therefore, in future studies MOFs might be one of the most powerful adsorbents for a green environment. Acknowledgements The authors would like to express their sincere thanks to Mr. Imteaz Ahmed, Enamul Haque, In Joong Kang, Jong Won Jun, Jaewoo Jeon and Miss Eun Young Park for their sincere efforts and contributions for this review article. This research was supported

by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (grant number 2012004528).

References
[1] B. Pawelec, R.M. Navarro, J.M. Campos-Martin, J.L.G. Fierro, Towards near zero-sulfur liquid fuels: a perspective review, Catal. Sci. Technol. 1 (2011) 2342. [2] R.N. Colvile, E.J. Hutchinson, J.S. Mindell, R.F. Warren, The transport sector as a source of air pollution, Atmos. Environ. 30 (2001) 15371565. [3] P.M. Ndegwaa, A.N. Hristov, J. Arogo, R.E. Shefeld, A review of ammonia emission mitigation techniques for concentrated animal feeding operations, Biosyst. Eng. 100 (2008) 453469. [4] G.W. Peterson, G.W. Wagner, A. Balboa, J. Mahle, T. Sewell, C.J. Karwacki, Ammonia vapor removal by Cu3 (BTC)2 and its characterization by MAS NMR, J. Phys. Chem. C 113 (2009) 1390613917. [5] C. Petit, T.J. Bandosz, Enhanced adsorption of ammonia on metal-organic framework/graphite oxide composites: analysis of surface interactions, Adv. Funct. Mater. 20 (2010) 111118. [6] G.W. Peterson, G.W. Wagner, A. Balboa, J. Mahle, T. Sewell, C.J. Karwack, Ammonia vapor removal by Cu3 (BTC)2 and its Characterization by MAS NMR, J. Phys. Chem. C 113 (2009) 1390613917. [7] D. Britt, D. Tranchemontagne, O.M. Yaghi, Metal-organic frameworks with high capacity and selectivity for harmful gases, Proc. Natl. Acad. Sci. USA 105 (2008) 1162321162. [8] C. Petit, C. Karwacki, G. Peterson, T.J. Bandosz, Interactions of ammonia with the surface of microporous carbon impregnated with transition metal chlorides, J. Phys. Chem. C 111 (2007) 1270512714. [9] A. Samokhvalov, B.J. Tatarchuk, Characterization of active sites, determination of mechanisms of H2 S, COS and CS2 sorption and regeneration of ZnO low-temperature sorbents: past, current and perspectives, Phys. Chem. Chem. Phys. 13 (2011) 31973209. [10] P. Kumar, C.-Y. Sung, O. Muraza, M. Cococcioni, S.A. Hashimi, A. McCormick, M. Tsapatsis, H2 S adsorption by Ag and Cu ion exchanged faujasites, Microporous Mesoporous Mater. 146 (2011) 127133. [11] H.F. Garces, H.M. Galindo, L.J. Garces, J. Hunt, A. Morey, S.L. Suib, Low temperature H2 S dry-desulfurization with zinc oxide, Microporous Mesoporous Mater. 127 (2010) 190197. [12] A.M. Vandenbroucke, R. Morent, N.D. Geyter, C. Leys, Non-thermal plasmas for non-catalytic and catalytic VOC abatement, J. Hazard. Mater. 195 (2011) 3054.

454

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456 [41] E. Haque, J.W. Jun, S.N. Talapaneni, A. Vinu, S.H. Jhung, Superior adsorption capacity of mesoporous carbon nitride with basic CN framework for phenol, J. Mater. Chem. 20 (2010) 1080110803. [42] Y. Wang, R.T. Yang, J.M. Heinzel, Desulfurization of jet fuel by -complexation adsorption with metal halides supported on MCM-41 and SBA-15 mesoporous materials, Chem. Eng. Sci. 63 (2008) 356365. [43] L. Zhang, W. Zhang, J. Shi, Z. Hua, Y. Li, J. Yan, A new thioether functionalized organic-inorganic mesoporous composite as a highly selective and capacious Hg2+ adsorbent, Chem. Commun. (2003) 210221. [44] X. Wang, T. Sun, J. Yang, L. Zhao, J. Jia, Low-temperature H2 S removal from gas streams with SBA-15 supported ZnO nanoparticles, Chem. Eng. J. 142 (2008) 4855. [45] J.-R. Li, R.J. Kuppler, H.-C. Zhou, Selective gas adsorption and separation in metal-organic frameworks, Chem. Soc. Rev. 38 (2009) 14771504. [46] H.-L. Jiang, Q. Xu, Porous metal-organic frameworks as platforms for functional applications, Chem. Commun. 47 (2011) 33513370. [47] C.-Y. Huang, M. Song, Z.-Y. Gu, H.-F. Wang, X.-P. Yan, Probing the adsorption characteristic of metal-organic framework MIL-101 for volatile organic compounds by quartz crystal microbalance, Environ. Sci. Technol. 45 (2011) 44904496. [48] J. Liu, P.K. Thallapally, B.P. McGrail, D.R. Brown, Progress in adsorption-based CO2 capture by metal-organic frameworks, J. Liu, Chem. Soc. Rev. 41 (2012) 23082322. [49] M. Seredych, E. Deliyanni, T.J. Bandosz, Role of microporosity and surface chemistry in adsorption of 4,6-dimethyldibenzothiophene on polymerderived activated carbons, Fuel 89 (2010) 14991507. [50] C.O. Ania, T.J. Bandosz, Importance of structural and chemical heterogeneity of activated carbon surfaces for adsorption of dibenzothiophene, Langmuir 21 (2005) 77527759. [51] M. Seredych, T.J. Bandosz, Removal of dibenzothiophenes from model diesel fuel on sulfur rich activated carbons, Appl. Catal. B Environ. 106 (2011) 133141. [52] A.U. Czaja, N. Trukhan, U. Mller, Enantioselective catalysis with homochiral metal-organic frameworks, Chem. Soc. Rev. 38 (2009), 1284-1256. [53] G. Frey, Hybrid porous solids: past, present, future, Chem. Soc. Rev. 37 (2008) 191214. [54] H. Li, M. Eddaoudi, M. OKeeffe, O.M. Yaghi, Design and synthesis of an exceptionally stable and highly porous metal-organic framework, Nature 402 (1999) 276279. [55] F. Schth, K.S.W. Sing, J. Weitkamp, Handbook of Porous Solids, John WileyVCH, Weinheim, 2002. [56] M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. OKeeffe, O.M. Yaghi, Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage, Science 295 (2002) 469472. [57] S.H. Jhung, N.A. Khan, Z. Hasan, Analogous porous metal-organic frameworks: synthesis, stability and application in adsorption, CrysEngComm 14 (2012) 70997109. [58] O.M. Yaghi, M. OKeeffe, N.W. Ockwig, H.K. Chae, M. Eddaoudi, J. Kim, Reticular synthesis and the design of new materials, Nature 423 (2003) 705714. [59] S. Kitagawa, R. Kitaura, S.-I. Noro, Functional porous coordination polymers, Angew. Chem. Int. Ed. 43 (2004) 23342375. [60] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.-H. Bae, J.R. Long, Carbon dioxide capture in metal-organic frameworks, Chem. Rev. 112 (2012) 724781. [61] M.P. Suh, H.J. Park, T.K. Prasad, D.-W. Lim, Hydrogen storage in metal-organic frameworks, Chem. Rev. 112 (2012) 782835. [62] H. Wu, Q. Gong, D.H. Olson, J. Li, Commensurate adsorption of hydrocarbons and alcohols in microporous metal organic frameworks, Chem. Rev. 112 (2012) 836868. [63] J.-R. Li, J. Sculley, H.-C. Zhou, Metal-organic frameworks for separations, Chem. Rev. 112 (2012) 869932. [64] P. Horcajada, R. Gref, T. Baati, P.K. Allan, G. Maurin, P. Couvreur, G. Frey, R.E. Morris, C. Serre, Metal-organic frameworks in biomedicine, Chem. Rev. 112 (2012) 12321268. [65] T. Uemura, N. Yanai, S. Kitagawa, Polymerization reactions in porous coordination polymers, Chem. Soc. Rev. 38 (2009) 12281236. [66] M. Kurmoo, Magnetic metal-organic frameworks, Chem. Soc. Rev. 38 (2009) 13531379. [67] J.Y. Lee, O.K. Farha, J. Roberts, K.A. Scheidt, S.T. Nguyen, J.T. Hupp, Metalorganic framework materials as catalysts, Chem. Soc. Rev. 38 (2009) 14501459. [68] J. Rocha, L.D. Carlos, F.A.A. Paz, D. Ananias, Luminescent multifunctional lanthanides-based metal-organic frameworks, Chem. Soc. Rev 40 (2011) 926940. [69] H. Furukawa, N. Ko, Y.B. Go, N. Aratani, S.B. Choi, Eunwoo Choi, A.. Yazaydin, R.Q. Snurr, M. OKeeffe, J. Kim, O.M. Yaghi, Ultrahigh porosity in metal-organic frameworks, Science 329 (2010) 424428. [70] A.R. Millward, O.M. Yaghi, Metal organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature, J. Am. Chem. Soc. 127 (2005) 1799817999. [71] Z. Zhao, X. Li, S. Huang, Q. Xia, Z. Li, Adsorption and diffusion of benzene on chromium-based metal organic framework MIL-101 synthesized by microwave irradiation, Ind. Eng. Chem. Res. 50 (2011) 22542261. [72] N.A. Khan, J.W. Jun, J.H. Jeong, S.H. Jhung, Remarkable adsorptive performance of a metal-organic framework, vanadium-benzenedicarboxylate (MIL-47), for benzothiophene, Chem. Commun. 47 (2011) 13061308.

[13] B. Guieysse, C. Hort, V. Platel, R. Munoz, M. Ondarts, S. Revah, Biological treatment of indoor air for VOC removal: potential and challenges, Biotechnol. Adv. 26 (2008) 398410. [14] J.-R. Li, Y. Ma, M.C. McCarthy, J. Sculley, J. Yub, H.-K. Jeong, P.B. Balbuena, H.-C. Zhou, Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks, Coord. Chem. Rev. 255 (2011) 17911823. [15] A. Samokhvalov, B.J. Tatarchuk, Review of experimental characterization of active sites and determination of molecular mechanisms of adsorption, desorption and regeneration of the deep and ultradeep desulfurization sorbents for liquid fuels, Catal. Rev. Sci. Eng. 52 (2010) 381410. [16] A. Stanislaus, A. Mara, M.S. Rana, Recent advances in the science and technology of ultra low sulfur diesel (ULSD) production, Catal. Today 153 (2010) 168. [17] S. Achmann, G. Hagen, M. Hmmerle, I. Malkowsky, C. Kiener, R. Moos, Sulfur removal from low-sulfur gasoline and diesel fuel by metal-organic frameworks, Chem. Eng. Technol. 33 (2010) 275280. [18] G. Crini, Non-conventional low-cost adsorbents for dye removal: a review, Bioresour. Technol. 97 (2006) 10611085. [19] A. Mittal, A. Malviya, D. Kaur, J. Mittal, L. Kurup, Studies on the adsorption kinetics and isotherms for the removal and recovery of methyl orange from wastewaters using waste materials, J. Hazard. Mater. 148 (2007) 229240. [20] S. Chen, J. Zhang, C. Zhang, Q. Yue, Y. Li, C. Li, Equilibrium and kinetic studies of methyl orange and methyl violet adsorption on activated carbon derived from phragmites australis, Desalination 252 (2010) 149156. [21] C.G. Daughton, T.A. Ternes, Pharmaceuticals and personal care products in the environment: agents of subtle change? Environ. Health Perspect. 107 (1999) 907938. [22] A.S. Mestre, J. Pires, J.M.F. Nogueira, A.P. Carvalho, Activated carbons for the adsorption of ibuprofen, Carbon 45 (2007) 19791988. [23] Z. Hasan, J. Jeon, S.H. Jhung, Adsorptive removal of naproxen and clobric acid from water using metal-organic frameworks, J. Hazard. Mater. 209210 (2012) 151157. [24] E.M. Cuerda-Correaa, J.R. Domnguez-Vargas, F.J. Olivares-Marn, J.B. Heredia, On the use of carbon blacks as potential low-cost adsorbents for the removal of non-steroidal anti-inammatory drugs from river water, J. Hazard. Mater. 177 (2010) 10461053. [25] J.K. Fawell, D. Sheahan, H.A. James, M. Hurst, S. Scott, Oestrogens and oestrogenic activity in raw and treated water in severn trent water, Water Res. 35 (2001) 12401244. [26] M. Carballa, F. Omil, J.M. Lema, Removal of pharmaceutical and personal care products (PPCPs) from municipal wastewaters by physico-chemical processes, Electron. J. Environ. Agric. Food Chem. 2 (2003) 309313. [27] C. Liu, Y. Huang, N. Naismith, J. Economy, Novel polymeric chelating bers for selective removal of mercury and cesium from water, Environ. Sci. Technol. 37 (2003) 42614268. [28] M.J. Manos, C.D. Malliakas, M.G. Kanatzidis, Heavy-metal-ion capture, ionexchange, and exceptional acid stability of the open-framework chalcogenide (NH4 )4In12 Se20 , Chem. Eur. J. 13 (2007) 5158. [29] B. Lee, Y. Kim, H. Lee, J. Yi, Synthesis of functionalized porous silicas via templating method as heavy metal ion adsorbents: the introduction of surface hydrophilicity onto the surface of adsorbents, Microporous Mesoporous Mater. 50 (2001) 7790. [30] F. Ke, L.-G. Qiu, Y.-P. Yuan, F.-M. Peng, X. Jiang, A.-J. Xie, Y.-H. Shen, J.-F. Zhu, Thiol-functionalization of metal-organic framework by a facile coordinationbased postsynthetic strategy and enhanced removal of Hg2+ from water, J. Hazard. Mater. 196 (2011) 3643. [31] B.-J. Zhu, X.-Y. Yu, Y. Jia, F.-M. Peng, B. Sun, M.-Y. Zhang, T. Luo, J.-H. Liu, X.-J. Huang, Iron 1,3,5-benzenetricarboxylic metal-organic coordination polymers prepared by solvothermal method and their application in effcient As(V) removal from aqueous solutions, J. Phys. Chem. C 116 (2012) 86018607. [32] I.V. Babich, J.A. Moulijn, Science and technology of novel processes for deep desulfurization of oil renery streams: a review, Fuel 82 (2003) 607631. [33] A. Dabrowski, P. Podko scielny, Z. Hubicki, M. Barczak, Adsorption of phenolic compounds by activated carbon a critical review, Chemosphere 58 (2005) 10491070. [34] K. Kadirvelu, K. Thamaraiselvi, C. Namasivayam, Removal of heavy metals from industrial wastewaters by adsorption onto activated carbon prepared from an agricultural solid waste, Bioresour. Technol. 76 (2001) 6365. [35] C.Y. Yin, M.K. Aroua, W.M. Ashri, W. Daud, Review of modications of activated carbon for enhancing contaminant uptakes from aqueous solutions, Sep. Purif. Technol. 52 (2007) 403415. [36] A. Zhou, X. Ma, C.S. Song, Liquid-phase adsorption of multi-ring thiophenic sulfur compounds on carbon materials with different surface properties, J. Phys. Chem. B 110 (2006) 46994707. [37] E. Deliyanni, M. Seredych, T.J. Bandosz, Interactions of 4,6dimethyldibenzothiophene with the surface of activated carbons, Langmuir 25 (2009) 93029312. [38] S. Velu, X. Ma, C. Song, Selective adsorption for removing sulfur from jet fuel over zeolite-based adsorbents, Ind. Eng. Chem. Res. 42 (2003) 52935304. [39] S. Choi, J.H. Drese, C.W. Jones, Adsorbent materials for carbon dioxide capture from large anthropogenic point sources, ChemSusChem 2 (2009) 796854. [40] J. Helminen, J. Helenius, E. Paatero, Adsorption equilibria of ammonia gas on inorganic and organic sorbents at 298.15 K, J. Chem. Eng. Data 46 (2001) 391399.

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456 [73] S.R. Caskey, A.G. Wong-Foy, A.J. Matzger, Dramatic tuning of carbon dioxide uptake via metal substitution in a coordination polymer with cylindrical pores, J. Am. Chem. Soc. 130 (2008) 1087010871. [74] T.G. Glover, G.W. Peterson, B.J. Schindler, D. Britt, O.M. Yaghi, MOF-74 building unit has a direct impact on toxic gas adsorption, Chem. Eng. Sci. 66 (2011) 163170. rey, G.D. Weireld, [75] L. Hamon, C. Serre, T. Devic, T. Loiseau, F. Millange, G. Fe Comparative study of hydrogen sulde adsorption in the MIL-53(Al, Cr, Fe), MIL-47(V), MIL-100(Cr), and MIL-101(Cr) metal organic frameworks at room temperature, J. Am. Chem. Soc. 131 (2009) 87758777. [76] J.R. Karra, K.S. Walton, Effect of open metal sites on adsorption of polar and nonpolar molecules in metal organic framework Cu-BTC, Langmuir 24 (2008) 86208626. [77] S.-H. Huo, X.-P. Yan, Metal-organic framework MIL-100(Fe) for the adsorption of malachite green from aqueous solution, J. Mater. Chem. 22 (2012) 74497455. [78] G. Blanco-Brieva, J.M. Campos-Martin, S.M. Al-Zahrani, J.L.G. Fierro, Effectiveness of metal-organic frameworks for removal of refractory organo-sulfur compound present in liquid fuels, Fuel 90 (2011) 190197. [79] B. Arstad, H. Fjellvg, K.O. Kongshaug, O. Swang, R. Blom, Amine functionalised metal organic frameworks (MOFs) as adsorbents for carbon dioxide, Adsorption 14 (2008) 755762. [80] E. Haque, J.E. Lee, I.T. Jang, Y.K. Hwang, J.-S. Chang, J. Jegal, S.H. Jhung, Adsorptive removal of methyl orange from aqueous solution with metal-organic frameworks, porous chromium-benzenedicarboxylates, J. Hazard. Mater. 181 (2010) 535542. [81] N.A. Khan, S.H. Jhung, Remarkable adsorption capacity of CuCl2 -loaded porous vanadium benzenedicarboxylate for benzothiophene, Angew. Chem. Int. Ed. 51 (2012) 11981201. [82] N.A. Khan, S.H. Jhung, Adsorptive removal of benzothiophene using porous copper-benzenetricarboxylate loaded with phosphotungstic acid, Fuel Process. Technol. 100 (2012) 4954. [83] I. Ahmed, Z. Hasan, N.A. Khan, S.H. Jhung, Adsorptive denitrogenation of fuels with porous metal-organic frameworks (MOFs): effect of acidity and basicity of MOFs, Appl. Catal. B: Environ. 129 (2013) 123129. [84] N.A. Khan, S.H. Jhung, Low-temperature loading of Cu+ species over porous metal-organic frameworks (MOFs) and adsorptive desulfurization with Cu+ loaded MOFs, J. Hazard. Mater. 237238 (2012) 180185. [85] L. Hamon, H. Leclerc, A. Ghou, L. Oliviero, A. Travert, J.-C. Lavalley, T. Devic, C. Serre, G. Ferey, G.D. Weireld, A. Vimont, G. Maurin, Molecular insight into the adsorption of H2 S in the exible MIL-53(Cr) and rigid MIL-47(V) MOFs: infrared spectroscopy combined to molecular simulations, J. Phys. Chem. C 115 (2011) 20472056. [86] V.C. Srivastava, An evaluation of desulfurization technologies for sulfur removal from liquid fuels, RSC Adv. 2 (2012) 759783. [87] F. Shi, M. Hammoud, L.T. Thompson, Selective adsorption of dibenzothiophene by functionalized metal organic framework sorbents, Appl. Catal. B Environ. 103 (2011) 261265. [88] G. Blanco-Brieva, J.M. Campos-Martin, S.M. Al-Zahrani, J.L.G. Fierro, Removal of refractory organic sulfur compounds in fossil fuels using MOF sorbents, Global NEST J. 12 (2010) 296304. [89] A.L. Nuzhdin, K.A. Kovalenko, D.N. Dybtsevb, G.A. Bukhtiyarova, Removal of nitrogen compounds from liquid hydrocarbon streams by selective sorption on metal-organic framework MIL-101, Mendeleev Commun. 20 (2010) 5758. [90] K.A. Cychosz, A.G. Wong-Foy, A.J. Matzger, Liquid phase adsorption by microporous coordination polymers: removal of organosulfur compounds, J. Am. Chem. Soc. 130 (2008) 69386939. [91] K.A. Cychosz, A.G. Wong-Foy, A.J. Matzger, Enabling cleaner fuels: desulfurization by adsorption to microporous coordination polymers, J. Am. Chem. Soc. 131 (2009) 1453814543. [92] S.S.-Y. Chui, S.M.-F. Lo, J.P.H. Charmant, A.G. Orpen, I.D. Williams, A chemically functionalizable nanoporous material [Cu3 (TMA)2 (H2O)3 ]n , Science 283 (1999) 11481150. [93] P.D.C. Dietzel, B. Panella, M. Hirscher, R. Blom, H. Fjellvag, Hydrogen adsorption in a nickel based coordination polymer with open metal sites in the cylindrical cavities of the desolvated framework, Chem. Commun. (2006) 959961. [94] D. Peralta, G. Chaplais, A. Simon-Masseron, K. Barthelet, G.D. Pirngruber, Metal-organic framework materials for desulfurization by adsorption, Energy Fuel 26 (2012) 49534960. [95] T.-H. Park, K.A. Cychosz, A.G. Wong-Foy, A. Dailly, A.J. Matzger, Gas and liquid phase adsorption in isostructural Cu3 [biaryltricarboxylate]2 microporous coordination polymers, Chem. Commun. 47 (2011) 14521454. [96] T. Loiseau, C. Serre, C. Huguenard, G. Fink, F. Taulelle, M. Henry, T. Bataille, G. Frey, A rationale for the large breathing of the porous aluminum terephthalate (MIL-53) upon hydration, Chem. Eur. J. 10 (2004) 13731382. [97] C. Serre, F. Millange, C. Thouvenot, M. Nogus, G. Marsolier, D. Lour, G. Frey, Very large breathing effect in the rst nanoporous chromium (III)-based solids: MIL-53 or CrIII (OH){O2 C C6 H4 CO2 }{HO2 C C6 H4 CO2 H}x H2 Oy , J. Am. Chem. Soc. 124 (2002) 1351913526. [98] K. Barthelet, J. Marrot, D. Riou, G. Frey, A breathing hybrid organic-inorganic solid with very large pores and high magnetic characteristics, Angew. Chem. Int. Ed. 41 (2002) 281284. [99] R.T. Yang, A.J. Hernandez-Maldonaldo, F.H. Yang, Desulfurization of transportation fuels with zeolites under ambient conditions, Science 301 (2003) 7981.

455

[100] Z.Y. Zhang, T.B. Shi, C.Z. Jia, W.J. Ji, Y. Chen, M.Y. He, Adsorptive removal of aromatic organosulfur compounds over the modied Na Y zeolites, Appl. Catal. B Environ. 82 (2008) 110. [101] A.J. Hernndez-Maldonado, G. Qi, R.T. Yang, Desulfurization of commercial fuels by -complexation: monolayer CuCl/-Al2 O3 , Appl. Catal. B Environ. 61 (2005) 212218. [102] X. Yang, L.E. Erickson, K.L. Hohn, Sol-Gel, Cu-Al2 O3 adsorbents for selective adsorption of thiophene out of hydrocarbon, Ind. Eng. Chem. Res. 45 (2006) 61696174. [103] Y. Wang, R.T. Yang, Desulfurization of jet fuel JP-5 light fraction by MCM-41 and SBA-15 supported cuprous oxide for fuel cell applications, Ind. Eng. Chem. Res. 48 (2009) 142147. [104] M. Jiang, F.T.T. Ng, Adsorption of benzothiophene on Y zeolites investigated by infrared spectroscopy and ow calorimetry, Catal. Today 116 (2006) 530536. [105] N.A. Khan, S.-M. Paek, S.H. Jhung, Effect of metal ions of analogous metalorganic frameworks (MOFs) on the adsorptive removal of benzothiophene from model fuel, in preparation. [106] M. Maes, M. Trekels, M. Boulhout, S. Schouteden, F. Vermoortele, L. Alaerts, D. Heurtaux, Y.-K. Seo, Y.K. Hwang, J.-S. Chang, I. Beurroies, R. Denoyel, K. Temst, A. Vantomme, P. Horcajada, C. Serre, D.E. De Vos, Selective removal of n-heterocyclic aromatic contaminants from fuels by Lewis acidic metalorganic frameworks, Angew. Chem. Int. Ed. 50 (2011) 42104214. [107] X.-X. Huang, L.-G. Qiu, W. Zhang, Y.-P. Yuan, X. Jiang, A.-J. Xie, Y.-H. Shena, J.F. Zhu, Hierarchically mesostructured MIL-101 metal-organic frameworks: supramolecular template-directed synthesis and accelerated adsorption kinetics for dye removal, CrystEngComm 14 (2012) 16131617. [108] E. Haque, J.W. Jun, S.H. Jhung, Adsorptive removal of methyl orange and methylene blue from aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235), J. Hazard. Mater. 185 (2011) 507511. [109] C. Chen, M. Zhang, Q. Guan, W. Li, Kinetic and thermodynamic studies on the adsorption of xylenol orange onto MIL-101(Cr), Chem. Eng. J. 183 (2012) 6067. [110] M. Maes, S. Schouteden, L. Alaerts, D. Depla, D.E. De Vos, Extracting organic contaminants from water using the metal-organic framework CrIII (OH){O2 C C6 H4 CO2 }, Phys. Chem. Chem. Phys. 13 (2011) 55875589. [111] A.C. Sudik, A.P. Ct, O.M. Yaghi, Metal-organic frameworks based on trigonal prismatic building blocks and the new acs topology, Inorg. Chem. 44 (2005) 29983000. [112] P.L. Llewellyn, P. Horcajada, G. Maurin, T. Devic, N. Rosenbach, S. Bourrelly, C. Serre, D. Vincent, S. Loera-Serna, Y. Filinchuk, G. Frey, Complex adsorption of short linear alkanes in the exible metal-organic-framework MIL-53(Fe), J. Am. Chem. Soc. 131 (2009) 1300213008. [113] M.M. Matlock, B.S. Howerton, K.R. Henke, D.A. Atwood, A pyridine-thiol ligand with multiple bonding sites for heavy metal precipitation, J. Hazard. Mater. 82 (2001) 5563. [114] I.H. Lee, Y.C. Kuan, J.M. Chern, Factorial experimental design for recovering heavy metals from sludge with ion-exchange resin, J. Hazard. Mater. 138 (2006) 549559. [115] R.S. Bai, E. Abraham, Studies on chromium (VI) adsorptiondesorption using immobilized fungal biomass, Bioresour. Technol. 87 (2003) 1726. O. Nivinskiene, I. Razmute, Copper, (II)-EDTA sorption onto chi[116] O. Gyliene, tosan and its regeneration applying electrolysis, J. Hazard. Mater. 137 (2006) 14301437. [117] H.J. Fan, P.R. Anderson, Copper and cadmium removal by Mn oxide-coated granular activated carbon, Sep. Purif. Technol. 45 (2005) 6167. [118] B. Yu, Y. Zhang, A. Shukla, S.S. Shukla, K.L. Dorris, The removal of heavy metal from aqueous solutions by sawdust adsorption-removal of copper, J. Hazard. Mater. B80 (2000) 3342. [119] Q.-R. Fang, D.-Q. Yuan, J. Sculley, J.-R. Li, Z.-B. Han, H.-C. Zhou, Functional mesoporous metal-organic frameworks for the capture of heavy metal ions and size-selective catalysis, Inorg. Chem. 49 (2010) 1163711642. [120] L. Mi, H. Hou, Z. Song, H. Han, Y. Fan, Polymeric zinc ferrocenyl sulfonate as a molecular aspirator for the removal of toxic metal ions, Chem. Eur. J. 14 (2008) 18141821. [121] L. Mi, H. Hou, Z. Song, H. Han, H. Xu, Y. Fan, S.-W. Ng, Rational construction of porous polymeric cadmium ferrocene-1,1 -disulfonates for transition metal ion exchange and sorption, Cryst. Growth Des. 7 (2007) 25532561. [122] S. Das, H. Kim, K. Kim, Metathesis in single crystal: complete and reversible exchange of metal ions constituting the frameworks of metal-organic frameworks, J. Am. Chem. Soc. 131 (2009) 38143815. [123] T.K. Prasad, D.H. Hong, M.P. Suh, High gas sorption and metal-ion exchange of microporous metal-organic frameworks with incorporated imide groups, Chem. Eur. J. 16 (2010) 1404314050. [124] C.A. Fernandez, P.K. Thallapally, R.K. Motkuri, S.K. Nune, J.C. Sumrak, J. Tian, J. Liu, Gas-Induced expansion and contraction of a uorinated metal-organic framework, Cryst. Growth Des. 10 (2010) 10371039. [125] C. Petit, T.J. Bandosz, Exploring the coordination chemistry of MOF-graphite oxide composites and their applications as adsorbents, Dalton Trans. 41 (2012) 40274035. [126] Q. Yang, S. Vaesen, M. Vishnuvarthan, F. Ragon, C. Serre, A. Vimont, M. Daturi, G.D. Weireld, G. Maurin, Probing the adsorption performance of the hybrid porous MIL-68(Al): a synergic combination of experimental and modelling tools, J. Mater. Chem. 22 (2012) 1021010220.

456

N.A. Khan et al. / Journal of Hazardous Materials 244245 (2013) 444456 [142] T.M. McDonald, D.M. DAlessandro, R. Krishnac, J.R. Long, Enhanced carbon dioxide capture upon incorporation of N,N -dimethylethylenediamine in the metal-organic framework CuBTTri, Chem. Sci. 2 (2011) 20222028. [143] S. Couck, J.F.M. Denayer, G.V. Baron, T. Rmy, J. Gascon, F. Kapteijn, An aminefunctionalized MIL-53 metal organic framework with large separation power for CO2 and CH4 , J. Am. Chem. Soc. 131 (2009) 63266327. [144] A. Demessence, D.M. DAlessandro, M.L. Foo, J.R. Long, Strong CO2 binding in a water-stable, triazolate-bridged metal-organic framework functionalized with ethylenediamine, J. Am. Chem. Soc. 131 (2009) 87848786. [145] J. An, S.J. Geib, N.L. Rosi, High and selective CO2 uptake in a cobalt adeninate metal organic framework exhibiting pyrimidine- and amino-decorated pores, J. Am. Chem. Soc. 132 (2010) 3839. [146] Z. Xiang, S. Leng, D. Cao, Functional group modication of metal-organic frameworks for CO2 capture, J. Phys. Chem. C 116 (2012) 1057310579. [147] J. Kim, S.-T. Yang, S.B. Choi, J. Sim, J. Kim, W.-S. Ahn, Control of catenation in CuTATB-n metal-organic frameworks by sonochemical synthesis and its effect on CO2 adsorption, J. Mater. Chem. 21 (2011) 30703076. [148] B. Xiao, P.S. Wheatley, X. Zhao, A.J. Fletcher, S. Fox, A.G. Rossi, I.L. Megson, S. Bordiga, L. Regli, K.M. Thomas, R.E. Morris, High-capacity hydrogen and nitric oxide adsorption and storage in a metal organic framework, J. Am. Chem. Soc. 129 (2007) 12031209. [149] B. Xiao, P.J. Byrne, P.S. Wheatley, D.S. Wragg, X. Zhao, A.J. Fletcher, K.M. Thomas, L. Peters, J.S.O. Evans, J.E. Warren, W. Zhou, R.E. Morris, Chemically blockable transformation and ultraselective low-pressure gas adsorption in a non-porous metal organic framework, Nat. Chem. 1 (2009) 289294. [150] M.J. Ingleson, R. Heck, J.A. Gould, M.J. Rosseinsky, Inorg. Chem. 48 (2009) 99869988. [151] S. Shimomura, M. Higuchi, R. Matsuda, K. Yoneda, Y. Hijikata, Y. Kubota, Y. Mita, J. Kim, M. Takata, S. Kitagawa, Selective sorption of oxygen and nitric oxide by an electron-donating exible porous coordination polymer, Nat. Chem. 2 (2010) 633637. [152] S.H. Jhung, J.-H. Lee, J.W. Yoon, C. Serre, G. Frey, J.-S. Chang, Microwave synthesis of chromium terephthalate MIL-101 and its benzene sorption ability, Adv. Mater. 19 (2007) 121124. [153] N.A. Khan, E. Haque, S.H. Jhung, Rapid syntheses of a metal-organic framework material Cu3 (BTC)2 (H2 O)3 under microwave: a quantitative analysis of accelerated syntheses, Phys. Chem. Chem. Phys. 12 (2010) 26252631. [154] E. Haque, N.A. Khan, J.H. Park, S.H. Jhung, Synthesis of a metal-organic framework material, iron terephthalate, by ultrasound, microwave, and conventional electric heating: a kinetic study, Chem. Eur. J. 16 (2010) 10461052. [155] E. Haque, N.A. Khan, C.M. Kim, S.H. Jhung, Syntheses of metal-organic frameworks and aluminophosphates under microwave heating: quantitative analysis of accelerations, Cryst. Growth Des. 11 (2011) 44134421. [156] J. Shi, Z. Zhao, Q. Xia, Y. Li, Z. Li, Adsorption and diffusion of ethyl acetate on the chromium-based metal-organic framework MIL-101, J. Chem. Eng. Data 56 (2011) 34193425. [157] Z. Zhao, X. Li, Z. Li, Adsorption equilibrium and kinetics of p-xylene on chromium-based metal organic framework MIL-101, Chem. Eng. J. 173 (2011) 150157.

[127] C. Petit, B. Levasseur, B. Mendoza, T.J. Bandosz, Reactive adsorption of acidic gases on MOF/graphite oxide composites, Microporous Mesoporous Mater. 154 (2012) 107112. [128] N. Heymans, S. Vaesen, G.D. Weireld, A complete procedure for acidic gas separation by adsorption on MIL-53 (Al), Microporous Mesoporous Mater. 154 (2012) 9399. [129] H. Dathe, E. Peringer, V. Roberts, A. Jentys, J.A. Lercher, Metal organic frameworks based on Cu2+ and benzene-1,3,5-tricarboxylate as host for SO2 trapping agents, C. R. Chimie 8 (2005) 753763. [130] D. Saha, Z. Bao, F. Jia, S. Deng, Adsorption of CO2 , CH4 , N2 O, and N2 on MOF-5, MOF-177 and zeolite 5A, Environ. Sci. Technol. 44 (2010) 18201826. [131] A.C. McKinlay, B. Xiao, D.S. Wragg, P.S. Wheatley, I.L. Megson, R.E. Morris, Exceptional behavior over the whole adsorption storage delivery cycle for NO in porous metal organic frameworks, J. Am. Chem. Soc. 130 (2008) 1044010444. [132] D.M. DAlessandro, B. Smit, J.R. Long, Carbon dioxide capture: prospects for new materials, Angew. Chem. Int. Ed. 49 (2010) 60586082. [133] Z. Liang, M. Marshall, A.L. Chaffee, CO2 adsorption-based separation by metal organic framework (Cu-BTC) versus zeolite (13X), Energy Fuel 23 (2009) 27852789. [134] P.L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G.D. Weireld, J.-S. Chang, D.-Y. Hong, Y.K. Hwang, S.H. Jhung, G. Frey, High uptakes of CO2 and CH4 in mesoporous metalsorganic frameworks MIL-100 and MIL101, Langmuir 24 (2008) 72457250. [135] A.. Yazaydn, A.I. Benin, S.A. Faheem, P. Jakubczak, J.J. Low, R.R. Willis, R.Q. Snurr, Enhanced CO2 adsorption in metal-organic frameworks via occupation of open-metal sites by coordinated water molecules, Chem. Mater. 21 (2009) 14251430. [136] P.D.C. Dietzel, V. Besikiotis, R. Blom, Application of metal-organic frameworks with coordinatively unsaturated metal sites in storage and separation of methane and carbon dioxide, J. Mater. Chem. 19 (2009) 73627370. [137] D. Britt, H. Furukawa, B. Wang, T.G. Glover, O.M. Yaghi, Highly efcient separation of carbon dioxide by a metal-organic framework replete with open metal sites, Proc. Natl. Acad. Sci. USA 106 (2009) 2063720640. [138] P.D.C. Dietzel, R.E. Johnsen, H. Fjellvag, S. Bordiga, E. Groppo, S. Chavanc, R. Blom, Adsorption properties and structure of CO2 adsorbed on open coordination sites of metal-organic framework Ni2 (dhtp) from gas adsorption, IR spectroscopy and X-ray diffraction, Chem. Commun. (2008) 51255127. [139] H.-Y. Cho, D.-A. Yang, J. Kim, S.-Y. Jeong, W.-S. Ahn, CO2 adsorption and catalytic application of Co-MOF-74 synthesized by microwave heating, Catal. Today 185 (2012) 3540. [140] D.-W. Jung, D.-A. Yang, J. Kim, J. Kim, W.-S. Ahn, Facile synthesis of MOF177 by a sonochemical method using 1-methyl-2-pyrrolidinone as a solvent, Dalton Trans. 39 (2010) 28832887. [141] D.-A. Yang, H.-Y. Cho, J. Kim, S.-T. Yang, W.-S. Ahn, CO2 capture and conversion using Mg-MOF-74 prepared by a sonochemical method, Energy Environ. Sci. 5 (2012) 64656473.

You might also like