Open access

Introductory Chapter: Metal Organic Frameworks (MOFs)

Written By

Eram Sharmin and Fahmina Zafar

Published: 12 October 2016

DOI: 10.5772/64797

From the Edited Volume

Metal-Organic Frameworks

Edited by Fahmina Zafar and Eram Sharmin

Chapter metrics overview

5,998 Chapter Downloads

View Full Metrics

1. Introduction

Over the past 50 decades, porous materials, from zeolites, coordination polymers to metal organic frameworks (MOFs), have gained considerable attention. The interesting feature is their porosity that allows the diffusion of guest molecules into the bulk structure. The shape and size of pores govern the shape and size selectivity of the guests to be incorporated. MOFs as defined by Yaghi et al. are porous structures constructed from the coordinative bonding between metal ions and organic linkers or bridging ligands (Figure 1) [1]. MOFs are formed by anchoring metal‐containing units or secondary‐building units (SBUs) with organic linkers, by coordination, yielding open frameworks that show exceptional feature of permanent porosity, stable framework, enormous surface area, and pore volume. The porosity is a consequence of long organic linkers that confer large storage space and numerous adsorption sites within MOFs. They also bear the ability to systematically vary and functionalize their pore structure [2, 3]. In the history of MOFs, a benchmark was represented by the synthesis of MOF‐5 (Zn4O(bdc)3, bdc = terephthalate) and HKUST‐1 (Cu3(btc)2, btc = 1,3,5‐benzenetricarboxylate) with high porosity and low pressure gas sorption, followed by the development of chromium(III) terephthalate (MIL‐101) with high chemical stability, MOF‐74 (Zn2(dhbdc), dhbdc = 2,5‐dihydroxy‐1,4‐benzenedicarboxylate) with low pressure adsorption of CO2, and several isostructural analogs of Mg‐MOF‐74 termed as IRMOF‐74‐I to IRMOF‐74‐XI, with large pore apertures to accommodate protein, NU‐110E with acetylene‐expanded hexatopic linker, having material highest experimental Brunauer‐Emmett‐Teller (BET) surface area of any porous material reported to date (7140 m2 g-1) Some examples of MOFs and their applications are given in Table 1 [115].

Figure 1.

Structure of MOF.

Application MOF  Metal  Ligand Year  Author Rf
Drug delivery MIL‐101 [Cr3O(OH,F,H2O)3(1,4‐bdc)3 and MIL‐100 Cr  1,4‐benzenedicarboxylate moieties (bdc) or H3btc: Benzene‐1,3,5‐tricarboxylate 2006 Patricia Horcajada et al. [4]
Methane
Storage
MOF‐5 Zn4(1,4‐bdc)3 Zn bdc 2002 Li and Eddaoudi, et al. [5, 6]
Adsorption
and storage
HKUST( Hong Kong University of Science and Technology)‐1 Cu2(H2O)2(CO2)4 Cu H3btc 2006 Rowsell and Yaghi [7]
Adsorption
and storage
IRMOF‐9 Zn4O(bpdc)3 Zn 4,4′‐biphenyldicarboxylate (bpdc) 2006 Rowsell and Yaghi [7]
Adsorption
and storage
MOF‐74, Zn2(C8H2O6) Zn 2,5‐dihydroxybenzene‐1,4‐dicarboxylic acid 2006 Rowsell and Yaghi [7]
(In) MIL‐68‐NH2 or IHM‐2 In bdc‐NH2: 2‐aminoterephthalates 2011 Savonnet and Farrusseng [8]
Drug
delivery
metal–organic Zn(bix) spheres with encapsulated DOX [DOX/Zn(bix)],
SN‐38 [SN‐38/Zn(bix)], CPT [CPT/Zn(bix)] and DAU [DAU/Zn(bix)] Doxorubicin (DOX), SN‐38, camptothecin (CPT) and daunomycin (DAU) 
Zn Bix: 1,4-bis(imidazol-1-ylmethyl)benzene 2010 Inhar Imaz et al. [9]
Soft
coupling–deprotection sequence
(In) MIL‐68‐NH‐ProFmoc and (In) MIL‐68‐NH‐Ala‐FmocIn *fluorenylmethyloxycarbonyl group (Fmoc), a base‐label protecting group for amines In Amino acid such as l‐proline (Pro‐OH) and D‐alanine (Ala‐OH) 2011 Jerome Canivet et al. [10]
Antibacterial Cu‐BTC(MOF‐199) Cu H3btc 2014 Rodrıguez et al. [11]
Highly potent bacteriocidal activity Co‐TDM Co H8 tdm: tetrakis [(3,5‐dicarboxyphenyl)‐oxamethyl] methane 2012 Wenjuan Zhuang et al. [12]
Delivery of nitric oxide MIL‐100(Fe or Cr)
and MIL‐127(Fe)
Fe, Cr or Fe tricarboxylate or tetracarboxylate 2014 Eubank et al. [13]
Antibacterial Ag2(O‐IPA)(H2O)·(H3O) and Ag5(PYDC)2(OH) Ag HO‐H2ipa = 5‐hydroxyisophthalic acid and H2pydc = pyridine‐3, 5‐dicarboxylic acid 2014 Xinyi Lu et al. [14]
Adsorption of CO2 over N2 Mn3(HCOO)6 ·DMF Mn 3‐nitrophthalic acid (H2npta) and 4,4′‐bipyridine (4,4′‐bipy) 2014 Ying‐Ping Zhao et al. [15]

Table 1.

Some examples of MOFs and their applications.

Advertisement

2. Chemistry

MOFs consist of both inorganic and organic units. The organic units (linkers/bridging ligands) consist of carboxylates, or anions, such as phosphonate, sulfonate, and heterocyclic compounds (Figures 2 and 3). The inorganic units are the metal ions or clusters termed as SBUs. Its geometry is determined by the coordination number, coordination geometry of the metal ions, and the nature of the functional groups. A variety of SBU geometries with different number of points of extension such as octahedron (six points), trigonal prism (six points), square paddle‐wheel (four points), and triangle (three points) have been observed in MOF structures (Figure 4). In principle, a bridging ligand (ditopic, tritopic, tetratopic, or multitopic linkers) reacts with a metal ion with more than one vacant or labile site. The final framework topology of MOF is governed by both SBU connectors and organic ligand linkers. Depending upon the nature of the system used, infinite‐extended polymeric or discrete‐closed oligomeric structures can arise (Figure 4). Metal‐containing units and organic linkers can be varied resulting in a variety of MOFs, tailored for different applications [3]. MOFs with large spaces may result in the formation of interpenetrating structures. Thus, it is very important to inhibit interpenetration by carefully choosing the organic linkers. The pore size is allowed to be tuned and spatial cavity arrangement be controlled, by judicious selection of metal centers and organic ligands and also by adjusting their conditions of synthesis. The large porosity allows their applications in adsorption and separation of gaseous molecules, catalysis, microelectronics, optics, sensing applications, bioreactors, drug delivery, and others. MOFs have pore openings up to 2‐nm size, which can accommodate small molecules. However, the pore openings rarely allow the inclusion of large molecules (e.g., proteins and enzymes). Attempts have been taken to increase the pore size to mesopore regime (pore size of 2–50 nm) and to decrease the crystal size to the nanometer scale. The large pore aperture benefits surface modification with a number of functionalities, without sacrificing the porosity of MOFs, also allowing the encapsulation of large molecule MOFs. The synthesis of MOFs involves reaction conditions and simple methods such as solvothermal, ionothermal, diffusion, microwave methods, ultrasound‐assisted, template‐directed syntheses, and others [2, 3].

An interesting and significant advancement in the field is to combine MOFs with functional nanoparticles, yielding new nanocomposite materials with unparalleled properties and performance. Nano‐MOFs are advantageous over conventional nanomedicines owing to their structural and chemical diversity, high loading capacity, and biodegradability. The final properties are dependent on the particle composition, size, and morphology. These can be obtained as either crystalline or amorphous materials. As soft porous crystals, framework flexibility (triggered by an external stimulus, e.g., mechanical stress, temperature, light interactions) may be shown by MOFs, also in the absence of guests or with no involvement of adsorption and desorption [13, 16].

Figure 2.

Some examples of organic ligands with carboxylic functionality used for the preparation of MOFs.

Figure 3.

Some examples of ligands containing nitrogen, sulfur, phosphorous and heterocycles used for the preparation of MOFs.

Figure 4.

MOFs resulting from different metal nodes and bridging ligands.

Advertisement

3. Metal biomolecule frameworks (BioMOFs)

Biomolecules are naturally and abundantly available. They are cost‐effective, rigid, and flexible with different coordination sites, rendering structurally diverse, biologically compatible MOFs. MOFs have also been synthesized from nontoxic endogenous cations (such as Ca, Mg, Fe, and Zn) and ligands consisting of naturally occurring derivatives or biomolecules [17]. These BioMOFs are usually biocompatible and suitable for biomedical applications [1747]. Such combinations of natural ligands with endogenous cations are also associated with several therapeutic effects (anti‐allergic, anti‐inflammatory, antimicrobial, anticarcinogenic activities). Table 2 shows some examples of BioMOFs and their applications [1847]. Such biologically and environmentally compatible MOFs are designed and constructed based on specific composition criteria governed by judiciously selecting metal ions and organic linkers as building blocks, which are nontoxic and biologically and environmentally compatible. Biomolecules such as amino acids, peptides, proteins, nucleobases, carbohydrates, and other natural products such as cyclodextrins, porphines, and some carboxylic acids (Figure 5) serve as emerging building blocks for the design and construction of metal‐biomolecule frameworks with novel and interesting properties and applications that cannot be obtained through the use of traditional organic linkers [17, 43, 44, 48, 49].

Aplication BioMOF Metal Ligand Year   AuthorRf
Ar and CH4 sorption [Cu(trans‐fum)] Cu Fum:Fumaric acid 2001 K. Seki et al
[18]
Reversible H2O
sorption/desorption
[Ni7(suc)6(OH)2
(H2O)2·2H2O
Ni Suc: Succinic
acid
2002 Forster et al.
[19]
[Ni7(suc)4(OH)6
(H2O)3]·7H2O
Ni Suc 2003 Guillou et al.
[20]
Sorption of more
than 30 kinds of
guests (e.g. DMF,
benzene,etc.);
structural change
[Mn3(HCOO)6]
·(CH3OH)
·(H2O)
Mn Formic acid 2004 Wang et al.
[21]
Selective CO2
and H2 sorption
Mn(HCOO)2·1/3
(C4H8O2)
Mn Formic
acid
2004 Dybtsev et al.
[22]
Adsorption Fe3O(MeOH)3(fum)3
(CO2CH3)]·4.5MeOH
Fe Fum 2004 Serre et al.
[23]
1,3‐Butanediol
sorption
[Ni2O(L‐Asp)
H2O]·4H2O
Ni Amino acid L‐Asp:l‐aspartic acid 2004 Anokhina et al.
[24]
Enantioselective
separation and
catalytic
Zn2(bdc)
(L‐lac)(DMF)
Zn bdc: 1,4‐
benzendicarboxylic
acid and L‐
lac:Lactic acid
2006 Dybtsev et al.
[25]
CO2 sorption [Ni2(L‐Asp)2
(4,4′‐bipy)]
·2H2O
Ni L‐Asp and
4,4′‐bipy : 1,2‐bis
(4‐pyridyl)ethane
2006 Vaidhyanathan
et al. [26]
H2 sorption Co2(L‐Asp)2
(4,4′‐bipy)]·2H2O
Co L‐Asp and 4,4′
‐bipy
2008 Zhu et al.
[27]
Heterogeneous
asymmetric
catalysts for
the methanolysis
of rac‐propylene
oxide
Ni2(L‐Asp)2
(4,4′‐bipy)
·(HCl)1.8(MeOH)
Ni L‐Asp and 4,
4′‐bipy
2008 Ingleson
et al. [28]
Heterogeneous
asymmetric
catalysts for
the methanolysis
of rac‐propylene
oxide
Cu2(l‐Asp)2
(bpe)·(HCl)2·
(H2O)2
Cu l‐Asp
and bpe: 1,2‐bis(4‐
pyridyl)ethane
2008 Ingleson
et al. [28]
Cation exchange
capabilities,
including
cationic drugs
and lanthanide
ions
Zn8(Ade)4(bpdc)6O·2
Me2NH2·
8DMF·11H2O
Zn Nucleobases
Adenine:Ade
and bpdc:
biphenyldicarboxylate
2009 An et al.
[29]
Selective
CO2. sorption
Co2(Ade)2(CO2CH3)2
·2DMF· 0.5H2O
Co Ade 2010 An et al.
[30]
Drug delivery
and imaging
Fe3O(MeOH)3(fumarate)3
(CO2CH3)]·4.5
MeOH and [Fe3O(MeOH)(C6H4O8)3Cl]·6MeOH
Fumarate and C6H4O8
is galactarate
2010 Horcajada
et al. [31]
Therapeutic agent BioMIL-1 Fe Nicotinic acid (pyridine-3-carboxylic acid, also called niacin or vitamin B3) 2010 Miller et al.
[32]
Reversible
flexible structure;
CO2, MeOH and
H2O sorption
[Zn(GlyAla)2]·(solvent) Zn Peptide,
Glycine‐adenine
2010 Rabone et al.
[33]
(γ‐CD)
(KOH)2
K Saccharides
γ ‐CD: cyclodextrins
2010 Smaldone et al.
[35]
Inclusion of
several molecules
(e.g. Rhodamine B,
4‐phenylazoplenol,etc.)
(γ‐CD)
(RbOH)2
Rb γ‐CD
γ‐CD is a (chiral) cyclic
oligosaccharide
composed of
eightR‐1,4‐
linkedD‐
glucopyranosyl
(R‐1,4‐D‐Glup)
2010 Smaldone et al.
[34]
Highly selective
adsorption of CO2
CD‐MOF‐2 Rb γ‐CD 2011 Jeremiah J.
Gassensmith et al.
[35]
Photostable O2
sensor
Zn8(Ade)4(bpdc)6·
O·2Me2NH2] loaded
with lanthanide
cations( Tb(III),
Sm(III), Eu(III)
and Yb(III))
Zn and
lanthanide
Ade and bpdc 2011 An et al.
[36]
M(II/III)
Gallates
Fe, Mn, Co
and Ni
H4gal: gallic
acid
2011 Saines et al.
[37]
Porous α‐CD‐MCF Rb α‐CD
R‐cyclodextrin
(R‐CD),
comprised of
sixR‐1,4‐
D‐Glupresidues
portrayed in
their stable
4C1 conformations
2012 Gassensmith
et al. [38]
Adsorption CD‐MOF‐1
and CD‐MOF‐2
CD‐MOF‐3
K, Rb and Cs γ‐CD 2012 Forgan et al.
[39]
Drug storage
and release
or for the
immobilization
and organization
of large biomolecules
Bio‐MOF‐100 Zn Ade 2012 Jihyun An
et al. [40]
MIL‐151 to ‐154 Zr H4gal 2014 Cooper et al.
[41]
Antibacterial BioMIL‐5 Zn AzA: azelaic
acid
2014 Tamames‐Tabar
et al. [42]
Antioxidant
carrier
Mg(H4gal) Mg H4gal 2015 Cooper et al.
[43]
Inclusion and
loading the
drug molecules
CD‐MOF‐1 Na β‐CD:
cyclodextrins
2015 Lu et al.
[44]
Electrochemical
nitrite
detection
MOF‐525 Zr H4tcpp: meso‐tetra
(4‐carboxyphenyl)
porphine
2015 Kung et al.
[45]
Ammonia uptake Al‐PMOF Al H4tcpp 2015 Wilcox et al.
[46]
Highly active
anti‐diabetic
activity
[Zn(ain)(atz)]n Zn Hatz : 5‐
aminotetrazole
and Hain: 2‐
amino‐4‐isonicotinic
2016 David Briones
et al. [47]

Table 2.

Some examples of BioMOFs and their applications.

Figure 5.

Examples of organic linkers used for the synthesis of BioMOFs.

Advertisement

4. Summary

MOFs find versatile applications as drug‐delivery agents, sensors, storage and separation systems, catalysts, and others. Nontoxic nano‐MOFs bearing tailored cores and surfaces can be used as nanodrug carriers for antitumor and anti‐HIV drugs (biomedicine, nontoxic, drug). MOFs with biomolecules as organic linkers are still in cradle stage in contrast to their counterparts bearing traditional organic linkers. However, biomolecules confer biological compatibility and easy recyclability to MOFs. They also confer unique characteristics such as chirality and specific recognition, self‐assembly characteristic, separation, ion exchange, and catalytic properties, also rendering bioinspired structures. In future, a better understanding and control of chemistry and design of MOFs may provide plethora of opportunities towards their structures, properties, and applications in different fields.

Advertisement

Acknowledgments

Dr Fahmina Zafar is thankful to UGC (New Delhi, India) for Dr DS Kothari Postdoctoral Fellowship, Ref. # F.4/2006(BSR)/13‐986/2013(BSR). The author is also thankful to Prof. Nahid Nishat (Mentor), Inorganic Materials Research Lab, Department of Chemistry, Jamia Millia Islamia (a Central University) New Delhi, India, for her kind support.

References

  1. 1. James SL. Metal‐organic frameworks. Chem. Soc. Rev. 2003;32:276–288. DOI: 10.1039/b200393g.
  2. 2. Furukawa H, Cordova KE, O’Keeffe M, Yaghi OM. The chemistry and applications of metal‐organic frameworks. Science 2013;341:1230444. DOI: 10.1126/science.12304.
  3. 3. Lu W, Wei Z, Gu Z‐Y, Liu T‐F, Park J, Park J, Tian J, Zhang M, Zhang Q, Gentle III T, Bosch M, Zhou H‐C. Tuning the structure and function of metal–organic frameworks via linker design. Chem. Soc. Rev. 2014;43:5561–5593. DOI: 10.1039/C4CS00003J.
  4. 4. Horcajada P, Serre C, Vallet‐Reg M, Sebban M, Taulelle F, Frey G. Metal–organic frameworks as efficient materials for drug delivery. Angew. Chem. Int. Ed. 2006;45:5974 –5978. DOI: 10.1002/anie.200601878.
  5. 5. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM. Design and synthesis of an exceptionally stable and highly porous metal‐organic framework. Nature 1999;402:276–279. doi:10.1038/46248.
  6. 6. Eddaoudi M, Kim J, Rosi N, Vodak D, Wachter J, O’Keefe M, Yaghi OM. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science 2002;295:469–472. DOI: 10.1126/science.1067208.
  7. 7. Rowsell JLC, Yaghi OM. Effects of functionalization, catenation, and variation of the metal oxide and organic linking units on the low‐pressure hydrogen adsorption properties of metal‐organic frameworks. J. Am. Chem. Soc. 2006;128:1304–1315. DOI: 10.1021/ja056639q.
  8. 8. Savonnet M, Farrusseng D. PCT Appl. WO2011048284, 2011.
  9. 9. Imaz I, Rubio‐Martınez M, Garcıa‐Fernandez L, Garcıa F, Ruiz‐Molina D, Hernando J, Puntes V, Maspoch D. Coordination polymer particles as potential drug delivery systems. Chem. Commun. 2010;46:4737–4739. DOI: 10.1039/c003084h.
  10. 10. Canivet J, Aguado S, Bergeret G, Farrusseng D. Amino acid functionalized metal–organic frameworks by a soft coupling–deprotection sequence. Chem. Commun. 2011;47:11650–11652. DOI: 10.1039/C1CC15541E.
  11. 11. Rodrıguez HS, Hinestroza JP, Ochoa‐Puentes C, Sierra CA, Soto CY. Antibacterial activity against Escherichia coli of Cu‐BTC (MOF‐199) metal‐organic framework immobilized onto cellulosic fibers. J. Appl. Polym. Sci. 2014. DOI: 10.1002/APP.40815.
  12. 12. Zhuang W, Yuan D, Li J‐R, Luo Z, Zhou H‐C, Bashir S, Liu J. Highly potent bactericidal activity of porous metal‐organic frameworks. Adv. Healthcare Mater. 2012;1:225–238. DOI: 10.1002/adhm.201100043.
  13. 13. Eubank JF, Wheatley PS, Lebars G, McKinlay AC, Leclerc H, Horcajada P, Daturi M, Vimont A, Morris RE, Serre C. Porous, rigid metal(III)‐carboxylate metal‐organic frameworks for the delivery of nitric oxide. Apl. Mater. 2014;2:124112. DOI: 10.1063/1.4902996.
  14. 14. Lu X, Ye J, Zhang D, Xie R, Bogale RF, Sun Y, Zhao LZQ, Ning G. Silver carboxylate metal‐organic frameworks with highly antibacterial activity and biocompatibility. J. Inorg. Biochem. 2014;138:114–121. DOI: 10.1016/j.jinorgbio.2014.05.005.
  15. 15. Zhao Y‐P, Yang H, Wang F, Du Z‐Y. A microporous manganese‐based metal‐organic framework for gas sorption and separation. J. Mol. Struct. 2014;1074:19–21. DOI: 10.1016/j.molstruc.2014.05.033.
  16. 16. Tamames‐Tabar C, Cunha D, Imbuluzqueta E, Ragon F, Serre C, Blanco‐Prieto M J, Horcajada P. Cytotoxicity of nanoscaled metal–organic frameworks J. Mater. Chem. B. 2014;2:262. DOI: 10.1039/c3tb20832j.
  17. 17. Imaz I, Rubio‐Martınez M, An J, Sole‐Font I, Rosi NL, Maspoch D. Metal–biomolecule frameworks (MBioFs). Chem. Commun. 2011;47:7287–7302. DOI: 10.1039/c1cc11202c.
  18. 18. Seki K, Takamizawa S, Mori W. Characterization of microporous copper (II) dicarboxylates (fumarate, terephthalate, and trans‐1,4‐cyclohexanedicarboxylate) by gas adsorption. Chem. Lett. 2001;122–123. doi:10.1246/cl.2001.122.
  19. 19. Forster PM, Cheetham AK. Open‐framework nickel succinate, [Ni7(C4H4O4)6(OH)2(H2O)2]2H2O: a new hybrid material with three‐dimensional ni‐o‐ni connectivity. Angew. Chem., Int. Ed. 2002;41:457–459N.
  20. 20. Guillou N, Livage C, van Beek W, Nogues M, Ferey G. A layered nickel succinate with unprecedented hexanickel units: structure elucidation from powder‐diffraction data, and magnetic and sorption properties. Angew. Chem. Int. Ed. 2003;42:643–647. DOI: 10.1002/anie.200390177.
  21. 21. Wang ZM, Zhang B, Fujiwara H, Kobayashi H, Kurmoo M. Mn3(HCOO)6: a 3D porous magnet of diamond framework with nodes of Mn‐centered MnMn4 tetrahedron and guest‐modulated ordering temperature. Chem. Commun. 2004;416–417. DOI: 0.1039/b314221c.
  22. 22. Dybtsev DN, Chun H, Yoon SH, Kim D, Kim K. Microporous manganese formate: a simple metal‐organic porous material with high framework stability and highly selective gas sorption properties. J. Am. Chem. Soc. 2004;126:32–33. DOI: 10.1021/ja038678c.
  23. 23. Serre C, Millange F, Surble S, Ferey G. A route to the synthesis of trivalent transition‐metal porous carboxylates with trimeric secondary building units. Angew. Chem. Int. Ed. 2004;43:6285–6289. DOI: 10.1002/anie.200454250.
  24. 24. Anokhina EV, Jacobson AJ. [Ni2O(L‐Asp)(H2O)2)].4H2)O: a homochiral 1D helical chain hybrid compound with extended Ni‐O‐Ni bonding. J. Am. Chem. Soc. 2004;126:3044–3045. DOI: 10.1021/ja031836f.
  25. 25. Dybtsev DN, Nuzhdin AL, Chun H, Bryliakov KP, Talsi EP, Fedin VP, Kim K. A homochiral metal–organic material with permanent porosity, enantioselective sorption properties, and catalytic activity. Angew. Chem. Int. Ed. 2006;45:916–920. DOI: 10.1002/anie.200503023.
  26. 26. Vaidhyanathan R, Bradshaw D, Rebilly JN, Barrio JP, Gould JA, Berry NG, Rosseinsky MJ. A family of nanoporous materials based on an amino acid backbone. Angew. Chem., Int. Ed. 2006;45:6495–6499. DOI: 10.1002/anie.200602242.
  27. 27. Zhu P, Gu W, Cheng FY, Liu X, Chen J, Yan SP, Liao DZ. Design of two 3D homochiral Co(II) metal–organic open frameworks by layered‐pillar strategy: structure and properties. Cryst. Eng. Commun. 2008;10: 963–967. DOI: 10.1039/B801177J.
  28. 28. Ingleson MJ, Barrio JP, Bacsa J, Dickinson C, Park H, Rosseinsky MJ. Generation of a solid Brønsted acid site in a chiral framework. Chem. Commun. 2008;11:1287–1289. DOI: 10.1039/B718443C.
  29. 29. An J, Geib SJ, Rosi NL. Cation‐triggered drug release from a porous zinc‐adeninate metal‐organic framework. J. Am. Chem. Soc. 2009;131:8376–8377. doi: 10.1021/ja902972w.
  30. 30. An J, Geib SJ, Rosi NL. High and selective CO2 uptake in a cobalt adeninate metal–organic framework exhibiting pyrimidine and amino‐decorated pores. J. Am. Chem. Soc. 2010;132:38–39. DOI: 10.1021/ja909169x.
  31. 31. Horcajada P, Chalati T, Serre C, Gillet B, Sebrie C, Baati T, Eubank JF, Heurtaux D, Clayette P, Kreuz C, Chang JS, Hwang YK, Marsaud V, Bories PN, Cynober L, Gil S, Ferey G, Couvreur P, Gref R. Porous metal–organic‐framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010;9:172–178. doi:10.1038/nmat2608.
  32. 32. Miller SR, Hertaux D, Baati T, Horcajada P, Greneche JM, Serre C. Biodegradable therapeutic MOFs for the delivery of bioactive molecules. Chem. Commun. 2010;46:4526–4528. DOI: 10.1039/C001181A.
  33. 33. Rabone J, Yue YF, Chong SY, Stylianou KC, Bacsa J, Bradshaw D, Darling GR, Berry NG, Khimyak YZ, Ganin AY, Wiper P, Claridge JB, Rosseinsky MJ. Science 2010;329:1053–1057. doi: 10.1126/science.1190672.
  34. 34. Smaldone RA, Forgan RS, Furukawa H, Gassensmith JJ, Slawin AMZ, Yaghi OM, Stoddart JF. Metal–organic frameworks from edible natural products. Angew. Chem. Int. Ed. 2010;49:8630–8634. DOI : 10.1002/anie.201002343.
  35. 35. Gassensmith JJ, Furukawa H, Smaldone RA, Forgan SS, Botros YY, Yaghi OM, Stoddart JF. J. Am. Chem. Soc. 2011;133:15312–15315. DOI: 10.1021/ja206525x.
  36. 36. An J, Shade CM, Chengelis‐Czegan DA, Petoud S, Rosi NL. Strong and reversible binding of carbon dioxide in a green metal organic framework. J. Am. Chem. Soc. 2011;133:1220–1223. DOI: 10.1021/ja206525x.
  37. 37. Saines PJ, Yeung HH‐M, Hester JR, Lennie AR, Cheetham AK. Detailed investigations of phase transitions and magnetic structure in Fe(III),Mn(II), Co(II) and Ni(II) 3,4,5‐trihydroxybenzoate (gallate) dihydrates by neutron and X‐ray diffraction. Dalton Trans. 2011;40:6401–6410. DOI: 10.1039/c0dt01687j.
  38. 38. Gassensmith JJ, Smaldone RA, Forgan RS, Wilmer CE, Cordes DB, Botros YY, Slawin AMZ, Snurr RQ, Stoddart JF. Polyporous metal‐coordination frameworks. Org. Lett. 2012;14:1460–1463. DOI: 10.1021/ol300199a.
  39. 39. Forgan RS, Smaldone RA, Gassensmith JJ, Furukawa H, Cordes DB, Li Q, Wilmer CE, Botros YY, Snurr RQ, Slawin AMZ, Stoddart JF. Nanoporous carbohydrate metalorganic frameworks. J. Am. Chem. Soc. 2012;134:406–417. DOI: 10.1021/ja208224f.
  40. 40. An J, Farha OK, Hupp JT, Pohl E, Yeh JI, Rosi NL. Metal‐adeninate vertices for the construction of an exceptionally porous metal‐organic framework. Nature Commun. 2012;3:604–609. DOI: 0.1038/ncomms1618.
  41. 41. Cooper L, Guillou N, Martineau C, Elkaim E, Taulelle F, Serre C, Devic T. ZrIV coordination polymers based on a naturally occurring phenolic derivative. Eur. J. Inorg. Chem. 2014;36:6281–6289. DOI:10.1002/ejic.201402891.
  42. 42. Tamames‐Tabar C, Imbuluzqueta E, Guillou N, Serre C, Miller SR, Elkaïm E, Horcajada P, Blanco‐Prieto M. J. A Zn azelate MOF: combining antibacterial effect. Cryst. Eng. Commun. 2015;17:456–462 DOI:10.1039/c4ce00885e.
  43. 43. Cooper L, Hidalgo T, Gorman M, Lozano‐Fernandez T, Simon‐Vazquez R, Olivier C, Guillou N, Serre C, Martineau C, Taulelle F, Damasceno‐Borges D, Maurin G, Gonzalez‐Fernandez A, Horcajada P, Devic T. A biocompatible porous Mg‐gallate metal‐organic framework as an antioxidant carrier. Chem. Commun. 2015;51:5848–5851. DOI: 10.1039/c5cc00745c.
  44. 44. Lu H, Yang X, Li S, Zhang Y, Sha J, Li C, Sun J. Study on a new cyclodextrin based metal–organic framework with chiral helices. Inorg. Chem. Commun. 2015;61:48–52. DOI: 10.1016/j.inoche.2015.08.015.
  45. 45. Kung C‐W, Chang T‐H, Chou L‐Y, Hupp JT, Farha OK, Ho K‐C. Porphyrin‐based metal‐organic framework thin films for electrochemical nitrite detection. Electrochem. Commun. 2015;58:51–56. DOI: 10.1016/j.elecom.2015.06.003.
  46. 46. Wilcox OT, Fateeva A, Katsoulidis AP, Smith MW, Stone CA, Rosseinsky MJ. Acid loaded porphyrin‐based metal–organic framework for ammonia uptake. Chem. Commun. 2015;51:14989–14991. DOI: 10.1039/C5CC06209H.
  47. 47. Briones D, Fernández B, Calahorro AJ, Fairen‐Jimenez D, Sanz R, Martínez F, Orcajo G, Sebastián ES, Seco JM, González CS, Llopis J, Rodríguez‐Diéguez A. Highly active anti‐diabetic metal‐organic framework. Cryst. Growth Des. 2016;16:537–540. DOI: 10.1021/acs.cgd.5b01274.
  48. 48. Keskin S, Kızılel S. Biomedical applications of metal organic frameworks. Ind. Eng. Chem. Res. 2011;50:1799–1812. DOI: 10.1021/ie101312k.
  49. 49. Imaz I, Rubio‐Martınez M, Garcıa‐Fernandez L, Garcıa F, Ruiz‐Molina D, Hernando J, Puntesa V, Maspoch D. Coordination polymer particles as potential drug delivery systems. Chem. Commun. 2010;46:4737–4739. DOI: 10.1039/C003084H.

Written By

Eram Sharmin and Fahmina Zafar

Published: 12 October 2016