Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 3293 word(s) 3293 2021-12-01 04:23:23 |
2 format correction Meta information modification 3293 2021-12-23 10:08:58 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Manousi, N.; Giannakoudakis, D.A.; Rosenberg, E. Extraction of Metal Ions with Metal–Organic Frameworks. Encyclopedia. Available online: https://encyclopedia.pub/entry/17494 (accessed on 27 July 2024).
Manousi N, Giannakoudakis DA, Rosenberg E. Extraction of Metal Ions with Metal–Organic Frameworks. Encyclopedia. Available at: https://encyclopedia.pub/entry/17494. Accessed July 27, 2024.
Manousi, Natalia, Dimitrios A. Giannakoudakis, Erwin Rosenberg. "Extraction of Metal Ions with Metal–Organic Frameworks" Encyclopedia, https://encyclopedia.pub/entry/17494 (accessed July 27, 2024).
Manousi, N., Giannakoudakis, D.A., & Rosenberg, E. (2021, December 23). Extraction of Metal Ions with Metal–Organic Frameworks. In Encyclopedia. https://encyclopedia.pub/entry/17494
Manousi, Natalia, et al. "Extraction of Metal Ions with Metal–Organic Frameworks." Encyclopedia. Web. 23 December, 2021.
Extraction of Metal Ions with Metal–Organic Frameworks
Edit

Metal–organic frameworks (MOFs) are crystalline porous materials composed of metal ions or clusters coordinated with organic linkers. Due to their extraordinary properties such as high porosity with homogeneous and tunable in size pores/cages, as well as high thermal and chemical stability, MOFs have gained attention in diverse analytical applications. 

MOFs metals extraction sample preparation microextraction spectrometry environmental samples food samples biological samples

1. Introduction

The terminology of metal–organic frameworks (MOFs) was initially introduced in 1995, when Yaghi and Li reported the synthesis of a new “zeolite-like” crystalline structure upon the polymeric coordination of Cu ions with 4,4′-bipyridine and nitrate ions, resulting to large rectangular channels [1]. MOFs are known to have superior characteristics, such as high surface area (theoretically up to 14.600 m2g−1) [2], porosity of uniform in structure and topology nanoscaled cavities, and satisfactory thermal and mechanical stability. Therefore, metal–organic frameworks were established as successful candidates for various applications like environmental remediation, detoxification media of toxic vapors, heterogeneous catalysis, gas storage, imaging and drug delivery, fuel cells, supercapacitors, and sensors [2][3][4][5][6][7][8][9][10][11][12][13].
In the field of analytical chemistry, MOFs have been employed in various analytical sample preparation methods including solid-phase extraction (SPE), dispersive solid-phase extraction (d-SPE), magnetic solid-phase extraction (MSPE), stir bar sorptive extraction (SBSE), and pipette tip solid-phase extraction (PT-SPE) [14][15][16][17][18]. Metal–organic frameworks have been also tested as stationary phases for high-performance liquid chromatography (HPLC), capillary electrochromatography (CEC), and gas chromatography (GC) with many advantages. Moreover, with the use of chiral MOFs, separation of chiral compounds has been also reported [19][20][21][22].
Metal–organic frameworks have been synthesized and successfully applied for the preconcentration of heavy metals from environmental samples prior to their detection/analysis with a spectroscopic technique. The most common metal ions used in MOFs are Zn(II), Cu(II), Fe(III), and Zr(IV), while terephthalic acid, trimesic acid, or 2-methylimidazole have been excessively used as organic linkers [23]. Many efforts have been made in order to overcome the low water stability of MOFs toward the preparation of suitable sorbents for the extraction of metal ions [24]. Compared with other sorbent materials, MOFs have a significant advantage of stable and homogeneous pores of specific sizes [25].
The effect of trace heavy metals on human health has attracted worldwide attention. Their increasing industrial, domestic, agricultural, and technological utilization has resulted in wide distribution in the environment. Metals such as cadmium, lead, mercury, chromium, and arsenic are considered as systemic toxicants and it, therefore, is essential to determine their levels in environmental samples [26]. Among the different analytical techniques that are widely used for the determination of metal ions are flame atomic absorption spectroscopy (FAAS), electrothermal atomic absorption spectroscopy (ETAAS), inductively coupled plasma optical emission spectrometry (ICP-OES), and inductively coupled plasma mass spectrometry (ICP-MS) [27][28][29].
Due to the low concentrations of metals and the presence of various interfering ions in complex matrices, the direct determination of such ions at trace levels is still challenging. Various novel materials including graphene oxide, activated carbon, carbon nanotubes, porous oxides, and metal–organic frameworks have been successfully employed for this purpose [30][31][32][33].

2. Stability of MOFs in Aquatic Environment

The stability of the framework in aqueous solutions depends on the strength of the metal–ligand coordination bonds [34]. The collapse of MOFs in the presence of water is linked to the competitive coordination of water and the organic linkers with the metal ions/nodes. The stability of the structure is also associated with other factors like the geometry of the coordination between metal-ligand, the surface hydrophobicity, the crystallinity, and the presence of defective sites [35]. The use of additives like graphite oxide, graphitic carbon nitride, nanoparticles, or the deposition on substrates such as carbon, fibers, or textiles, can have a positive effect on the framework stability [36][37][38][39][40][41][42]. In order to evaluate the stability and as a result the properness of utilizing a MOF for adsorption application, the pH and the temperature under which the preconcentration of the metal will take place, must be considered.
The strength of the coordination between the organic moieties and the metal ions can be described in general according to the HSAB (hard/soft acid/base) principles [9][42]. Zr4+, Fe3+, Cr3+, and Al3+ are regarded as hard acidic metal ions, while Cu2+, Zn2+, Ni2+, Mn2+, and Ag+ as soft ones [34]. On the other hand, carboxylate-based linkers act as hard bases, while azolate ligands (such as pyrazolates, triazolates, or imidazolates) as soft bases. For that reason, most of the Zr-based UiO (University of Oslo) and MIL-53(Fe) (Material Institut Lavoisier) series possess remarkable water stability, while for instance one of the most known and studied MOF, HKUST-1 (Hong Kong University of Science and Technology) does not. On representative paradigm of Zn-based water-stable structure is the zeolitic imidazolate framework (ZIF), formed from imidazolate ligands and Zn2+.
When used in analytical chemistry, MOFs must be stable both under adsorption and under desorption conditions. Usually, adsorption of metal ions takes place under weakly acidic conditions (pH = 5–6), while desorption is performed predominately with the addition of a strong acid. However, even though many MOFs are stable under adsorption conditions, they are decomposed with the addition of strong acids like nitric, hydrochloric, and sulfuric acid [24][28]. Other reagents that have been employed for the elution of metal ions without decomposing the MOF material are ethylenediaminetetraacetic acid (EDTA), sodium chloride (NaCl), or sodium hydroxide (NaOH) solution in EDTA or in thiourea.

3. Mechanisms of Metal Ions Extraction with Metal–Organic Frameworks

MOFs, as well as their composites, have been successfully applied as adsorbents for various heavy metal/metalloid species. The adsorption of the latter from aquatic environments is still among the ultimate research targets, and there are plenty of reports in which adsorption/removal of heavy metals was a success story [43][44][45]. Although, not all MOFs are water-stable as discussed above. The most widely reported interactions/mechanisms are collected in Figure 1 [46]. In many cases, more than one mechanism is responsible for the high adsorptive capability of MOFs. The binding/interaction sites can be either the metal or the clusters as well as the linkers. In order to enhance the adsorptive capability and/or selectivity, the functionalization of the linkers, with groups as hydroxyl, thiol, or amide, is a well-explored and successive strategy.
Figure 1. A schematic illustration of the interactions/mechanisms involved in the adsorption of metals by metal–organic frameworks (MOFs).
Lewis acid–base interactions are the most common adsorption mechanism of metal ions by metal–organic frameworks [47]. The presence of O-, S-, and N-containing groups that act as Lewis bases is very important for the preconcentration of the various ionic species from aqueous solution since metal ions act as Lewis acids. The donor atoms of the MOFs are present in the molecules of the organic linkers. Pre- or post-synthesis functionalization of the frameworks can increase the number of O-, S-, or N-containing groups in order to enhance the adsorption selectivity and efficiency of the target metal ions. Since Lewis acid–base interactions are critical for metal adsorption onto the donor atoms of the MOFs, it is obvious that the pH of the solution plays the most critical role, influencing the adsorption process and kinetics. In low pH value, those atoms are protonated, and adsorption cannot take place due to the repulsive forces of the cationic form of metal with the positively charged adsorption sites [48]. However, by increasing the pH of the aqueous samples that contain the metal ions, the donor atoms of the adsorbent are deprotonated and they become favorable for complex formation and sorption of the target analytes. In basic solutions, the addition of hydroxide may lead to complex formation and precipitation of many metals, therefore, after a certain pH value, any further increase can lead to a decrease of the sorption efficiency [49][50].
Adsorption by coordination is another adsorption mechanism in which the functionalization plays a key role. For instance, Liu et al. showed that the post-synthetic modification of Cr-MIL-101 with incorporation of -SH functionalities led to an improvement of Hg(II) removal, even at ultra-low concentrations [51]. This improvement was linked to the coordination between Hg(II) with the -SH groups. The incorporation of thiol-containing benzene-1,4-dicarboxylic acid (BDC) linkers in the case of UiO-66 MOF resulted in a material capable of simultaneously adsorbing As(III) and As(V) oxyanions. The adsorption of the former occurred via coordination to the -SH groups, while of the latter by the binding of the oxyanions to the Zr6O4(OH)4 cluster via hydroxyl exchange [52]. The hydroxyl exchange mechanism was also proposed as the predominant capturing pathway in the study of Howard and co-workers [53], in which they studied the adsorption of Se(IV) and Se(VI) in water by seven Zr-based MOFs (UiO-66, UiO-66-NH2, UiO-66-(NH2)2, UiO-66-(OH)2, UiO-67, NU-1000, and NU-1000BA).
Additionally, the adsorption mechanism with metal–organic frameworks can be enhanced via the chelation mechanism, after functionalization of MOFs with compounds that can form chelating complexes with the metal ions [54]. For example, functionalization of metal–organic frameworks with dithizone can enhance Pb extraction by forming penta-heterocycle chelating complex compounds. In this case, the binding sites of the chelating molecules are also protonated in low pH values and adsorption cannot take place. Adsorption capacity increases with increasing pH until a certain point, normally at a pH value of 5 to 6. Further increase in pH value can lead to precipitation of the target analytes, due to hydrolysis [55].
In the case of the physical-based adsorption, various interactions can be responsible for the elevated adsorptive capability of MOFs as mentioned above. The net charge of the framework and the presence of specific functional groups have a positive impact on the extent of the physical interactions [56]. The manipulation of the above can be achieved by grafting of particular species/groups into the framework or by tuning the net charge as a result of the solution pH in which the adsorption takes place.
The electrostatic interactions between the negatively charged adsorption sites of MOFs with the oppositely charged adsorbates are the most widely reported pathway [57]. The diffusion of the metal ions toward the active sites prior to the blockage of the outer entrances of the channels is also an important aspect and so, the volume, geometry, and size of the pores are of paramount importance [58].

4. Sample Preparation Techniques for the Extraction of Metal Ions

Solid-phase extraction (SPE) is a well-established analytical technique that has been widely used for the extraction, preconcentration, clean-up, and class fractionation of various pollutants from environmental, biological, and food samples. Different sorbents have been evaluated for the SPE procedure usually placed into cartridges [59]. MOFs have been employed as sorbents for the solid-phase extraction. In a typical SPE application, the sorbent is conditioned to increase the effective surface area and to minimize potential interferences, prior to the loading of the sample solution onto a solid-phase [60][61][62]. The analytes are retained onto the active sites of the sorbent and the undesired components are washed out. Finally, elution of the analytes with the desired solvent is carried out [49].
SPE and other conventional sample preparation techniques like protein precipitation and liquid–liquid extraction (LLE) have fundamental drawbacks such as time-consuming complex steps, difficulty in automation, and need for large amounts of sample and organic solvents. Novel extraction techniques, including MSPE, d-SPE, SBSE, and PT-SPE, have been developed in order to overcome these problems. Figure 2 shows the typical steps of MSPE and d-SPE. Recently, MOFs have been used as sorbents for these extraction techniques [63].
Figure 2. Typical magnetic solid-phase extraction (MSPE) and dispersive solid-phase extraction (d-SPE) procedures for the enrichment and analysis of trace metal ions.
Dispersive solid-phase extraction is performed by direct addition of the sorbent into the solution that contains the target analytes. Various MOF materials have been employed for the d-SPE of metal ions from complex sample matrices. After a certain time, the sorbent is retrieved from the solution with centrifugation or filtration and the solution is discarded. Elution with an appropriate solvent is performed and the liquid phase is isolated for instrumental analysis. The dispersion is often enhanced by stirring, vortex mixing, or ultrasound irradiation, in order to enable an efficient transfer of the target analytes to the active sites of the sorbent. Therefore, several devices including shakers, vortex mixers, and ultrasonic probes and baths have been implemented for sorbent dispersion. Until today, the ultrasound-assisted dispersive solid-phase microextraction is the most common d-SPE approach [24][64].
MSPE is based on the use of sorbents with magnetic properties. There are several different procedures to fabricate magnetic MOFs that have been employed to prepare sorbents for MSPE. The most common approaches are the direct post-synthesis of magnetic MOF materials with magnetic nanoparticles and the second one, in situ growth of magnetic nanoparticles during the synthesis of the framework. In the first case, the desired MOF and the magnetic nanoparticles (Fe3O4) are synthesized separately and mixed under sonication. For the in situ approach, the MOF is added to a solution containing the reagents for the synthesis of Fe3O4 in order to give a magnetic material. Moreover, single-step MOF coating can take place by adding the Fe3O4 nanoparticles into a mixture of inorganic and organic precursors for MOF synthesis. Carbonization of some MOFs can shape magnetic nanoparticles due to aggregation of the metallic component of the MOF. At the same time, the organic linker is converted to a porous carbon. Finally, the layer-by-layer approach is based on the sequential immobilization of the different components of the MOFs into a functionalized support.
For the typical MSPE procedure, a magnetic sorbent is added to the sample for sufficient time in order to ensure a quantitative extraction. After this period of time, an external magnet is employed to retrieve the sorbent and the sample is discarded. The sorbent is washed and an appropriate solvent is added in order to desorb the analytes. After magnetic separation, the eluent can be directly analyzed or it can be evaporated and reconstitute in an appropriate solvent prior to the analysis [65][66].
Other extraction techniques that can be coupled with MOFs in order to extract different analytes from complex matrices are stir bar sorptive extraction (SBSE) and pipette tip solid-phase extraction (PT-SPE). SBSE is an equilibrium technique, initially introduced by Baltussen et al. In this technique, extraction of the analytes takes place onto the surface of a coated stir bar [67][68][69]. PT-SPE is a miniaturized form of SPE in which ordinary pipette tips act as the extracting column and small amount of sorbent is packed inside the tip [70][71]. Only a small range of SBSE and PT-SPE sorbents are commercially available, which limits the possible applications of those techniques. MOF materials have been successfully used as coatings for stir bars and as packed sorbents in pipette tips [67][68][69][70][71].
Although MOFs pose several benefits as extraction sorbents for SPE, MSPE d-SPE, SBSE, and PT-SPE, their water stability and selectivity have to be enhanced with appropriate functional groups or pore functionalization. Therefore, the type of metal–organic framework and the possible functionalization should be carefully chosen. Other parameters that should be thoroughly investigated are the pH value of the sample solution, the extraction and desorption time, the desorption solvent, etc.
As mentioned before, the pH of the sample solution is one of the most critical parameters for the extraction of heavy metals from aqueous samples. Therefore, the pH value has to be optimized carefully in order to allow the Lewis acid–base interactions between the sorbent and the target analytes and to prevent precipitation due to hydrolysis.
The mass of the MOF material, as well as the extraction time, are other parameters that can influence the extraction step and require optimization. First of all, an optimum adsorbent amount is necessary in order to maximize the extraction efficiency. Certain extraction time is also required to facilitate the interaction between the analytes and adsorption sites of the MOF material. Finally, the sample volume and the volume of the eluent has to be optimized in order to provide a higher enrichment factor that is possible.
Regarding the desorption step, among the parameters that should be thoroughly investigated are the type, the volume, and the concentration of the eluent. In most cases, elution can be achieved with acidic solutions of nitric or hydrochloric acid. The presence of H+ ions weakens the interaction between the analyte and the MOF, as it competes for binding with the active sites of the adsorbent. However, decomposition of most MOFs has been observed in acidic conditions. Other reagents that have been used for the elution of metal ions without decomposing the MOF material are EDTA, NaCl, NaOH in EDTA, NaOH in thiourea, etc. Furthermore, enough desorption time should be provided in order to enable the quantitative elution of the adsorbed analytes.
Other parameters that can be investigated are the stirring speed, salt addition, the use of ultrasonic radiation, etc., depending on the extraction procedure [69][70][71][72][73]. The optimization of the experimental parameters can be performed by evaluating one-factor-at-a-time or by performing Design of Experiments (DoE), such as Box–Behnken experimental design [74].
Finally, the effect of potentially interfering ions that naturally occur in the various sample matrices, the adsorption capacity of the MOF material, as well as the reusability of the sorbent should be also evaluated [69][70][71][72][73].

5. Applications of Metal–Organic Frameworks for the Extraction of Metal Ions

The applications of MOFs for the extraction of metal ions from environmental, biological, and food samples, as well as the obtained recoveries and limits of detection (LODs), are summarized in Table 1.
Table 1. Applications of metal–organic frameworks for the extraction of metal ions.

Analyte

Organic Linker of MOF

Metal of MOF

Modification

Matrix

Sample Preparation Technique

Detection Technique

Recovery (%)

LOD

(ng mL−1)

Reusability

Ref.

Pd(II)

Trimesic acid

Cu

Fe3O4@Py

Fish, sediment, soil, water,

MSPE

FAAS

96.8–102.5

0.37

-

[75]

 

Malonic acid

Ag

-

Water

SPE

FAAS

>95

0.5

Up to 5 times

[60]

Pb(II)

Trimesic Acid

Cu

DHz, Fe3O4

Water

MSPE

ETAAS

97–102

0.0046

At least 80 times

[76]

Trimesic Acid

Cu

Fe3O4@SH

Rice, pig liver, tea, water

MSPE

FAAS

>95

0.29–0.97

-

[72]

meso-tetra(4-carboxyphenyl) porphyrin

Zr

-

Cereal, beverage, water

d-SPE

FAAS

90–107

1.78

Up to 42 times

[77]

 

Trimesic acid

Cu

Fe3O4@4-(5)-imidazole-dithiocarboxylic acid

Fish, canned tune

MSPE

CVAAS

95–102

10

At least 12 times

[78]

Hg(II)

Trimesic acid

Cu

Thiol-modified silica

Fish, sediment, water

d-SPE

CV-AAS

91–102

0.02

-

[73]

3′5,5′-azobenzenetetracarboxylic acid

Cu

-

Tea, mushrooms

d-SPE

AFS

Average 93.3

>0.58 mg kg−1

Up to 3 times

[79]

Benzoic acid and meso-tetrakis(4-Carboxyphenyl)porphyrin

Zr

-

Fish

PT-SPE

CVAAS

74.3–98.7

20 × 10−3

At least 15 times

[71]

Cu (II)

Aminoterephthalic acid

Zn

Fe3O4

Water

MSPE

ETAAS

98–102

0.073

 

[28]

Cd(II)

Terephthalic acid

Fe

Fe3O4@MAA, AMSA

Water

MSPE

FAAS

>96

0.04

Up to 10 times

[80]

Th(IV)

2 –hydroxyterephthalic acid

Zr

-

Water

d-SPE

Spectrophotometry

>90

0.35

At least 25 times

[81]

[1,1′-biphenyl]-4-carboxylic acid

Eu

-

Water

Probe

UV

N.A.

24.2

N.A.

[82]

U(VI)

4,4′,4″-(1,3,5-triazine-2,4,6-triyltriimino)tris-benzoic acid

Te

-

Water

d-SPE

ICP-MS

94.2–98.0

0.9

At least 3 times

[83]

Se(IV), Se(VI)

Terephthalic acid

Cr

Fe3O4@dithiocarbamate

Water, agricultural samples

MSPE

ETAAS

>92

0.01

Up to 12 times

[84]

Cd(II), Pb(II)

Trimesic acid

Cu

Fe3O4@Py

Fish, sediment

water

MSPE

FAAS

92.0–103.3

0.2–1.1

-

[85]

Cd(II) Pb(II) Ni(II)

Trimesic acid

Cu

Fe3O4@TAR

Sea food, agricultural samples

MSPE

FAAS

83–112

0.15–0.8

-

[86]

Cd(II), Pb(II), Zn(II) Cr(III)

Trimesic acid

Cu

Fe3O4-benzoyl isothiocyanate

Vegetables

MSPE

FAAS

80–114

0.12–0.7

-

[49]

Terephthalic acid

Fe

Fe3O4-ethylenediamine

Agricultural samples

MSPE

FAAS

87.3–110

0.15–0.8

-

[87]

Cd(II), Pb(II), Ni(II), Zn(II)

Trimesic Acid

Cu

Fe3O4@DHz

Fish, sediment, soil, water

MSPE

FAAS

88–104

0.12–1.2

-

[55]

Pb(II), Cu(II)

Trimesic acid

Dy

-

Water

d-SPE

FAAS

95–105

0.26–0.40

At least 5 times

[50]

Cd(II), Co(II), Cr(III), Cu(II), Pb(II)

4-bpmb

Zn

-

Water

d-SPE

ICP-OES

90–110

0.01–1

-

[24]

Co(II), Cu(II), Pb(II), Cd(II), Ni(II), Cr(III), Mn(II)

4,4′-oxybisbenzoic acid

Cd

Fe3O4

Water

MSPE

ICP-OES

>90

0.3–1

-

[88]

Hg(II), Cr(VI) Pb(II) Cd(II)

Terephthalic acid

Cu

Dithioglycol

Tea

d-SPE

AFS, AAS

95–99

Not mentioned

Up to 3 times

[89]

References

  1. Yaghi, O.; Li, H. Hydrothermal synthesis of a Metal-Organic Framework containing large rectangular channels. J. Am. Chem. Soc. 1995, 117, 10401–10402.
  2. Farha, O.K.; Eryazici, I.; Jeong, N.C.; Hauser, B.G.; Wilmer, C.E.; Sarjeant, A.A.; Snurr, R.Q.; Nguyen, S.T.; Yazaydin, A.Ö.; Hupp, J.T. Metal-Organic Framework materials with ultrahigh surface areas: Is the sky the limit? J. Am. Chem. Soc. 2012, 134, 15016–15021.
  3. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of Metal-Organic Frameworks. Science 2013, 341, 1230444.
  4. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndta, K.; Pastréa, J. Metal-Organic Frameworks—Prospective industrial applications. J. Mater. Chem. 2006, 16, 626–636.
  5. Taylor-Pashow, K.; Della Rocca, J.; Xie, Z.; Tran, S.; Lin, W. Postsynthetic modifications of iron-carboxylate nanoscale Metal–Organic Frameworks for imaging and drug delivery. J. Am. Chem. Soc. 2009, 131, 14261–14263.
  6. Corma, A.; Garcia, H.; Llabres i Xamena, F.X.L.I. Engineering Metal Organic Frameworks for heterogeneous catalysis. Chem. Rev. 2010, 110, 4606–4655.
  7. Getman, R.; Bae, Y.; Wilmer, C.; Snurr, R. Review and analysis of molecular simulations of methane, hydrogen, and acetylene storage in Metal–Organic Frameworks. Chem. Rev. 2011, 112, 703–723.
  8. Lu, K.; Aung, T.; Guo, N.; Weichselbaum, R.; Lin, W. Nanoscale Metal-Organic Frameworks for Therapeutic, Imaging, and Sensing Applications. Adv. Mater. 2018, 30, 1707634.
  9. Dhakshinamoorthy, A.; Asiri, A.; Garcia, H. 2D Metal–Organic Frameworks as multifunctional materials in heterogeneous catalysis and electro/photocatalysis. Adv. Mater. 2019, 31, 1900617.
  10. Giannakoudakis, D.A.; Bandosz, T.J. Detoxification of Chemical Warfare Agents, 1st ed.; Springer International Publishing: Cham, Switzerland, 2018.
  11. Hashemi, B.; Zohrabi, P.; Raza, N.; Kim, K. Metal-Organic Frameworks as advanced sorbents for the extraction and determination of pollutants from environmental, biological, and food media. TrAC Trends Anal. Chem. 2017, 97, 65–82.
  12. Raza, W.; Kukkar, D.; Saulat, H.; Raza, N.; Azam, M.; Mehmood, A.; Kim, K. Metal-Organic Frameworks as an emerging tool for sensing various targets in aqueous and biological media. TrAC Trends Anal. Chem. 2019, 120, 115654.
  13. DeCoste, J.B.; Peterson, G.W. Metal-organic frameworks for air purification of toxic chemicals. Chem. Rev. 2014, 114, 5695–5727.
  14. Wang, G.; Lei, Y.; Song, H. Evaluation of Fe3O4@SiO2–MOF-177 as an advantageous adsorbent for magnetic solid-phase extraction of phenols in environmental water samples. Anal. Methods 2014, 6, 7842–7847.
  15. Dai, X.; Jia, X.; Zhao, P.; Wang, T.; Wang, J.; Huang, P.; He, L.; Hou, X. A combined experimental/computational study on metal-organic framework MIL-101(Cr) as a SPE sorbent for the determination of sulphonamides in environmental water samples coupling with UPLC-MS/MS. Talanta 2016, 15, 581–588.
  16. Hu, C.; He, M.; Chen, B.; Zhong, C.; Hu, B. Polydimethylsiloxane/metal-organic frameworks coated stir bar sorptive extraction coupled to high performance liquid chromatography-ultraviolet detector for the determination of estrogens in environmental water samples. J. Chromatogr. A 2013, 1310, 21–30.
  17. Lv, Z.; Sun, Z.; Song, C.; Lu, S.; Chen, G.; You, J. Sensitive and background-free determination of thiols from wastewater samples by MOF-5 extraction coupled with high-performance liquid chromatography with fluorescence detection using a novel fluorescence probe of carbazole-9-ethyl-2-maleimide. Talanta 2016, 161, 228–237.
  18. Yan, Z.; Wu, M.; Hu, B.; Yao, M.; Zhang, L.; Lu, Q.; Pang, J. Electrospun UiO-66/polyacrylonitrile nanofibers as efficient sorbent for pipette tip solid-phase extraction of phytohormones in vegetable samples. J. Chromatogr. A 2018, 1542, 19–27.
  19. Rocío-Bautista, P.; Taima-Mancera, T.; Pasán, J.; Pino, V. Metal-Organic Frameworks in green analytical chemistry. Separations 2019, 6, 33.
  20. Yusuf, K.; Aqel, A.; Alothman, Z. Metal-Organic Frameworks in chromatography. J. Chromatogr. A 2014, 1348, 1–16.
  21. Fei, Z.; Zhang, M.; Zhang, J.; Yuan, L. Chiral metal–organic framework used as stationary phases for capillary electrochromatography. Anal. Chim. Acta 2014, 830, 49–55.
  22. González-Rodríguez, G.; Taima-Mancera, I.; Lago, A.; Ayala, J.; Pasán, J.; Pino, V. Mixed functionalization of organic ligands in UiO-66: A tool to design Metal–Organic Frameworks for tailored microextraction. Molecules 2019, 24, 3656.
  23. Rocío-Bautista, P.; González-Hernández, P.; Pino, V.; Pasán, J.; Afonso, A. Metal-Organic Frameworks as novel sorbents in dispersive-based microextraction approaches. TrAC Trends Anal. Chem. 2017, 90, 114–134.
  24. Tahmasebi, E.; Masoomi, M.; Yamini, Y.; Morsali, A. Application of mechanosynthesized azine-decorated Zinc(II) Metal–Organic Frameworks for highly efficient removal and extraction of some heavy-metal ions from aqueous samples: A comparative study. Inorg. Chem. 2014, 54, 425–433.
  25. Rocío-Bautista, P.; Pacheco-Fernández, I.; Pasán, J.; Pino, V. Are Metal-Organic Frameworks able to provide a new generation of solid-phase microextraction coatings?—A review. Anal. Chim. Acta 2016, 939, 26–41.
  26. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy metals toxicity and the environment. EXS 2012, 101, 133–164.
  27. Anthemidis, A.; Kazantzi, V.; Samanidou, V.; Kabir, A.; Furton, K. An automated flow injection system for metal determination by flame atomic absorption spectrometry involving on-line fabric disk sorptive extraction technique. Talanta 2016, 156, 64–70.
  28. Wang, Y.; Xie, J.; Wu, Y.; Hu, X. A magnetic Metal-Organic Framework as a new sorbent for solid-phase extraction of copper(II), and its determination by electrothermal AAS. Microchim. Acta 2014, 181, 949–956.
  29. Samanidou, V.; Sarakatsianos, I.; Manousi, N.; Georgantelis, D.; Goula, A.; Adamopoulos, K. Detection of mechanically deboned meat in cold cuts by inductively coupled plasma-mass spectrometry. Pak. J. Anal. Environ. Chem. 2018, 19, 115–121.
  30. Zhang, Y.; Zhong, C.; Zhang, Q.; Chen, B.; He, M.; Hu, B. Graphene oxide–TiO2 composite as a novel adsorbent for the preconcentration of heavy metals and rare earth elements in environmental samples followed by on-line inductively coupled plasma optical emission spectrometry detection. RSC Adv. 2015, 5, 5996–6005.
  31. Narin, I.; Soylak, M.; Elçi, L.; Doğan, M. Determination of trace metal ions by AAS in natural water samples after preconcentration of pyrocatechol violet complexes on an activated carbon column. Talanta 2000, 52, 1041–1046.
  32. Sitko, R.; Zawisza, B.; Malicka, E. Modification of carbon nanotubes for preconcentration, separation and determination of trace-metal ions. TrAC Trends Anal. Chem. 2012, 37, 22–31.
  33. Manousi, N.; Zachariadis, G.; Deliyanni, E.; Samanidou, V. Applications of Metal-Organic Frameworks in food sample preparation. Molecules 2018, 23, 2896.
  34. Feng, M.; Zhang, P.; Zhou, H.; Sharma, V.K. Water-Stable Metal-Organic Frameworks for aqueous removal of heavy metals and radionuclides: A review. Chemosphere 2018, 209, 783–800.
  35. Zuluaga, S.; Fuentes-Fernandez, E.M.A.; Tan, K.; Xu, F.; Li, J.; Chabal, Y.J.; Thonhauser, T. Understanding and controlling water stability of MOF-74. J. Mater. Chem. A. 2016, 4, 5176–5183.
  36. Giannakoudakis, D.A.; Bandosz, T.J. Defectous UiO-66 MOF Nanocomposites as Reactive Media of Superior Protection against Toxic Vapors. ACS Appl. Mater. Interfaces 2019, in press.
  37. Giannakoudakis, D.A.; Hu, Y.; Florent, M.; Bandosz, T.J. Smart textiles of MOF/g-C3N4 nanospheres for the rapid detection/detoxification of chemical warfare agents. Nanoscale Horiz. 2017, 2, 356–364.
  38. Giannakoudakis, D.A.; Travlou, N.A.; Secor, J.; Bandosz, T.J. Oxidized g-C3N4 Nanospheres as Catalytically Photoactive Linkers in MOF/g-C3N4 Composite of Hierarchical Pore Structure. Small 2017, 13, 1601758.
  39. Xiang, W.; Zhang, Y.; Lin, H.; Liu, C.J. Nanoparticle/Metal–Organic Framework Composites for Catalytic Applications: Current Status and Perspective. Molecules 2017, 2, 2103.
  40. Ahmed, I.; Jhung, S.H. Composites of metal-organic frameworks: Preparation and application in adsorption. Mater. Today 2014, 17, 136–146.
  41. Petit, C.; Bandosz, T.J. Engineering the surface of a new class of adsorbents: Metal-organic framework/graphite oxide composites. J. Coll. Interface Sci. 2015, 447, 139–151.
  42. Alfarra, A.; Frackowiak, E.; Beguin, F. The HSAB concept as a means to interpret the adsorption of metal ions onto activated carbons. Appl. Surf. Sci. 2004, 228, 84–92.
  43. Howarth, A.J.; Liu, Y.; Hupp, J.; Farha, O.K. Metal–Organic Frameworks for applications in remediation of oxyanion/cation-contaminated water. CrystEngComm 2015, 17, 7245–7253.
  44. Li, S.; Chen, Y.; Pei, X.; Zhang, S.; Feng, X.; Zhou, J.; Wang, B. Water purification: Adsorption over Metal-Organic Frameworks. Chin. J. Chem. 2016, 34, 175–185.
  45. Hasan, Z.; Jhung, S.H. Removal of hazardous organics from water using Metal-Organic Frameworks (MOFs): Plausible mechanisms for selective adsorptions. J. Hazard. Mater. 2015, 283, 329–339.
  46. Khan, N.A.; Hasan, Z.; Jhung, S.H. Adsorptive removal of hazardous materials using Metal-Organic Frameworks (MOFs): A review. J. Hazard. Mater. 2013, 244, 444–456.
  47. Vu, T.A.; Le, G.H.; Dao, C.D.; Dang, L.Q.; Nguyen, K.T.; Nguyen, Q.K.; Dang, P.T.; Tran, H.T.K.; Duong, Q.T.; Nguyen, T.V.; et al. Arsenic removal from aqueous solutions by adsorption using novel MIL-53(Fe) as a highly efficient adsorbent. RSC Adv. 2015, 5, 5261–5268.
  48. Zhang, J.; Xiong, Z.; Li, C.; Wu, C. Exploring a thiol-functionalized MOF for elimination of lead and cadmium from aqueous solution. J. Mol. Liq. 2016, 221, 43–50.
  49. Hassanpour, A.; Hosseinzadeh-Khanmiri, R.; Babazadeh, M.; Abolhasani, J.; Ghorbani-Kalhor, E. Determination of heavy metal ions in vegetable samples using a magnetic Metal–Organic Framework nanocomposite sorbent. Food Addit. Contam. Part. A 2015, 32, 725–736.
  50. Jamali, A.; Tehrani, A.; Shemirani, F.; Morsali, A. Lanthanide Metal–Organic Frameworks as selective microporous materials for adsorption of heavy metal ions. Dalton Trans. 2016, 45, 9193–9200.
  51. Ke, F.; Qiu, L.G.; Yuan, Y.P.; Peng, F.M.; Jiang, X.; Xie, A.J.; Shen, Y.H.; Zhu, J.F. Thiol-Functionalization of Metal-Organic Framework by a facile coordination-based postsynthetic strategy and enhanced removal of Hg2+ from water. J. Hazard. Mater. 2011, 196, 36–43.
  52. Audu, C.O.; Nguyen, H.G.T.; Chang, C.; Katz, M.J.; Mao, L.; Farha, O.K.; Hupp, J.T.; Nguyen, S.T. The dual capture of AsV and AsIII by UiO-66 and analogues. Chem. Sci. 2016, 7, 6492–6498.
  53. Howarth, A.J.; Katz, M.J.; Wang, T.C.; Platero-Prats, A.E.; Chapman, K.W.; Hupp, J.T.; Farha, O.K. High efficiency adsorption and removal of selenate and selenite from water using Metal-Organic Frameworks. J. Am. Chem. Soc. 2015, 137, 7488–7494.
  54. Fang, Q.-R.; Yuan, D.-Q.; Sculley, J.; Li, J.-R.; Han, Z.-B.; Zhou, H.-C. Functional mesoporous metal-organic frameworks for the capture of heavy metal ions and size-selective catalysis. Inorg. Chem. 2010, 49, 11637–11642.
  55. Taghizadeh, M.; Asgharinezhad, A.; Pooladi, M.; Barzin, M.; Abbaszadeh, A.; Tadjarodi, A. A novel magnetic Metal-Organic Framework nanocomposite for extraction and preconcentration of heavy metal ions, and its optimization via experimental design methodology. Microchim. Acta 2013, 180, 1073–1084.
  56. Liu, B.; Jian, M.P.; Liu, R.P.; Yao, J.F.; Zhang, X.W. Highly efficient removal of arsenic(III) from aqueous solution by zeolitic imidazolate frameworks with different morphology. Coll. Surf. A Physicochem. Eng. Asp. 2015, 481, 358–366.
  57. Rahimi, E.; Mohaghegh, N. Removal of toxic metal ions from sungun acid rock drainage using mordenite zeolite, graphene nanosheets, and a Novel Metal–Organic Framework. Mine Water Environ. 2015, 35, 18–28.
  58. Jian, M.P.; Liu, B.; Zhang, G.S.; Liu, R.P.; Zhang, X.W. Adsorptive removal of arsenic from aqueous solution by zeolitic imidazolate framework-8 (ZIF-8) nanoparticles. Coll. Surf. A Physicochem. Eng. Asp. 2014, 465, 67–76.
  59. Andrade-Eiroa, A.; Canle, M.; Leroy-Cancellieri, V.; Cerdà, V. Solid-Phase extraction of organic compounds: A critical review (Part I). TrAC Trends Anal. Chem. 2016, 80, 641–654.
  60. Salarian, M.; Ghanbarpour, A.; Behbahani, M.; Bagheri, S.; Bagheri, A. A Metal-Organic Framework sustained by a nanosized Ag12 cuboctahedral node for solid-phase extraction of ultra traces of lead(II) ions. Microchim. Acta 2014, 181, 999–1007.
  61. Li, X.; Xing, J.; Chang, C.; Wang, X.; Bai, Y.; Yan, X.; Liu, H. Solid-Phase extraction with the Metal-Organic Framework MIL-101(Cr) combined with direct analysis in real time mass spectrometry for the fast analysis of triazine herbicides. J. Sep. Sci 2014, 37, 1489–1495.
  62. Chatzimichalakis, P.F.; Samanidou, V.F.; Verpoorte, R.; Papadoyannis, I.N. Development of a validated HPLC method for the determination of B-complex vitamins in pharmaceuticals and biological fluids after solid phase extraction. J. Sep. Sci. 2004, 27, 1181–1188.
  63. Manousi, N.; Raber, G.; Papadoyannis, I. Recent advances in microextraction techniques of antipsychotics in biological fluids prior to liquid chromatography analysis. Separations 2017, 4, 18.
  64. Ghorbani, M.; Aghamohammadhassan, M.; Chamsaz, M.; Akhlaghi, H.; Pedramrad, T. Dispersive solid-phase microextraction. TrAC Trends Anal. Chem. 2019, 118, 793–809.
  65. Giakisikli, G.; Anthemidis, A.N. Magnetic materials as sorbents for metal/metalloid preconcentration and/or separation. A review. Anal. Chim. Acta 2013, 789, 1–16.
  66. Maya, F.; Cabello, C.P.; Frizzarin, R.M.; Estela, J.M.; Palomino, G.T.; Cerdà, V.; Turnes, G. Magnetic solid-phase extraction using Metal-Organic Frameworks (MOFs) and their derived carbons. TrAC Trends Anal. Chem. 2017, 90, 142–152.
  67. Baltussen, E.; Sandra, P.; David, F.; Cramers, C. Stir bar sorptive extraction (SBSE), a novel extraction technique for aqueous samples: Theory and principles. J. Microcolumn Sep. 1999, 11, 737–747.
  68. Hu, C.; He, M.; Chen, B.; Zhong, C.; Hu, B. Sorptive extraction using polydimethylsiloxane/metal–organic framework coated stir bars coupled with high performance liquid chromatography-fluorescence detection for the determination of polycyclic aromatic hydrocarbons in environmental water samples. J. Chromatogr. A 2014, 1356, 45–53.
  69. Xiao, Z.; He, M.; Chen, B.; Hu, B. Polydimethylsiloxane/metal-organic frameworks coated stir bar sorptive extraction coupled to gas chromatography-flame photometric detection for the determination of organophosphorus pesticides in environmental water samples. Talanta 2016, 156, 126–133.
  70. Hashemi, S.H.; Kaykhaii, M.; Keikha, A.J.; Mirmoradzehi, E.; Sargazi, G. Application of response surface methodology for optimization of metal-organic framework based pipette-tip solid phase extraction of organic dyes from seawater and their determination with HPLC. BMC Chem. 2019, 13, 59.
  71. Rezaei Kahkha, M.; Daliran, S.; Oveisi, A.; Kaykhaii, M.; Sepehri, Z. The mesoporous porphyrinic zirconium Metal-Organic Framework for pipette-tip solid-phase extraction of mercury from fish samples followed by cold vapor atomic absorption spectrometric determination. Food Anal. Methods 2017, 10, 2175–2184.
  72. Wang, Y.; Chen, H.; Tang, J.; Ye, G.; Ge, H.; Hu, X. Preparation of magnetic Metal-Organic Frameworks adsorbent modified with mercapto groups for the extraction and analysis of lead in food samples by flame atomic absorption spectrometry. Food Chem. 2015, 181, 191–197.
  73. Sohrabi, M. Preconcentration of mercury(II) using a thiol-functionalized Metal-Organic Framework nanocomposite as a sorbent. Microchim. Acta 2013, 181, 435–444.
  74. Box, G.E.P.; Behnken, D.W. Some new three level designs for the study of quantitative variables. Technometrics 1960, 2, 455–475.
  75. Bagheri, A.; Taghizadeh, M.; Behbahani, M.; Akbar Asgharinezhad, A.; Salarian, M.; Dehghani, A.; Ebrahimzadeh, H.; Amini, M. Synthesis and characterization of magnetic Metal-Organic Framework (MOF) as a novel sorbent, and its optimization by experimental design methodology for determination of palladium in environmental samples. Talanta 2012, 99, 132–139.
  76. Wang, Y.; Xie, J.; Wu, Y.; Ge, H.; Hu, X. Preparation of a functionalized magnetic Metal–Organic Framework sorbent for the extraction of lead prior to electrothermal atomic absorption spectrometer analysis. J. Mater. Chem. A 2013, 1, 8782–8789.
  77. Tokalıoglu, S.; Yavuz, E.; Demir, S.; Patat, S. Zirconium-Based highly porous Metal-Organic Framework (MOF-545) as an efficient adsorbent for vortex assisted-solid-phase extraction of lead from cereal, beverage and water samples. Food Chem. 2017, 237, 707–715.
  78. Tadjarodi, A.; Abbaszadeh, A. A magnetic nanocomposite prepared from chelator-modified magnetite (Fe3O4) and HKUST-1 (MOF-199) for separation and preconcentration of mercury(II). Microchim. Acta 2016, 183, 1391–1399.
  79. Wu, Y.; Xu, G.; Wei, F.; Song, Q.; Tang, T.; Wang, X.; Hu, Q. Determination of Hg (II) in tea and mushroom samples based on Metal-Organic Frameworks as solid-phase extraction sorbents. Microporous Mesoporous Mater. 2016, 235, 204–210.
  80. Moradi, S.E.; Dadfarnia, S.; Emami, S.; Shabani, A.M.H. Sulfonated metal organic framework loaded on iron oxide nanoparticles as a new sorbent for the magnetic solid phase extraction of cadmium from environmental water samples. Anal. Methods 2016, 8, 6337–6346.
  81. Moghaddam, Z.; Kaykhaii, M.; Khajeh, M.; Oveisi, A. Synthesis of UiO-66-OH zirconium Metal-Organic Framework and its application for selective extraction and trace determination of thorium in water samples by spectrophotometry. Spectrochim. Acta A 2018, 194, 76–82.
  82. Liu, W.; Dai, X.; Wang, Y.; Song, L.; Zhang, L.-J.; Zhang, D.; Xie, J.; Chen, L.; Diwu, J.; Wang, J.; et al. Ratiometric monitoring of thorium contamination in natural water using a dual-emission luminescent europium organic framework. Environ. Sci. Technol. 2018, 53, 332–341.
  83. Liu, W.; Dai, X.; Bai, Z.; Wang, Y.; Yang, Z.; Zhang, L.; Xu, L.; Chen, L.; Li, Y.; Gui, D.; et al. Highly sensitive and selective uranium detection in natural water systems using a luminescent mesoporous Metal–Organic Framework equipped with abundant lewis basic sites: A combined batch, X-ray absorption spectroscopy, and first principles simulation investigation. Environ. Sci. Technol 2017, 51, 3911–3921.
  84. Kalantari, H.; Manoochehri, M. A nanocomposite consisting of MIL-101(Cr) and functionalized magnetite nanoparticles for extraction and determination of selenium(IV) and selenium(VI). Microchim. Acta 2018, 185, 196.
  85. Sohrabi, M.; Matbouie, Z.; Asgharinezhad, A.; Dehghani, A. Solid-Phase extraction of Cd(II) and Pb(II) using a magnetic metal-organic framework, and their determination by FAAS. Microchim. Acta 2013, 180, 589–597.
  86. Ghorbani-Kalhor, E. A metal-organic framework nanocomposite made from functionalized magnetite nanoparticles and HKUST-1 (MOF-199) for preconcentration of Cd(II), Pb(II), and Ni(II). Microchim. Acta 2016, 183, 2639–2647.
  87. Babazadeh, M.; Hosseinzadeh-Khanmiri, R.; Abolhasani, J.; Ghorbani-Kalhor, E.; Hassanpour, A. Solid-phase extraction of heavy metal ions from agricultural samples with the aid of a novel functionalized magnetic metal–organic framework. RSC Adv. 2015, 5, 19884–19892.
  88. Safari, M.; Yamini, Y.; Masoomi, M.; Morsali, A.; Mani-Varnosfaderani, A. Magnetic Metal-Organic Frameworks for the extraction of trace amounts of heavy metal ions prior to their determination by ICP-AES. Microchim. Acta 2017, 184, 1555–1564.
  89. Wu, Y.; Xu, G.; Liu, W.; Yang, J.; Wei, F.; Li, L.; Zhang, W.; Hu, Q. Postsynthetic modification of copper terephthalate metal-organic frameworks and their new application in preparation of samples containing heavy metal ions. Microporous Mesoporous Mater. 2015, 210, 110–115.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 1.1K
Revisions: 2 times (View History)
Update Date: 23 Dec 2021
1000/1000
Video Production Service