Anda di halaman 1dari 11

Universidad San Francisco de Quito

César Zambrano, PhD

Fundamentals of Inorganic Chemistry

Dynamics in Metal-Organic Frameworks (MOFs). The Materials Beyond: Complex arrays for
complex applications

Juan Esteban Zurita


00125756

December 9th, 2018

1
Abstract
Metal-Organic Frameworks are a type of organometallic complexes that are constructed by joining inorganic metal
containing units with organic linkers through a technique called reticular synthesis, that yields highly symmetrical, multiple
purpose materials that have been characterizes extensively by chemists in many areas of research, like nanomaterials,
computation, gas storage, separation processes, molecular machines, macromolecular confinement and many more.
Chemists today look for ways in which to understand and dominate the necessary variables involved in the synthesis of
complex solids to create rational, predictable and potentially intelligent materials. Metal-Organic Frameworks have
achieved many goals in the chemistry of porous materials, as having the greatest surface areas known for synthetic
materials, and also as having the potential of being the basis of models to design complexity and dynamics within rigid and
stable polymers. The authors claim that there are infinite possibilities for the creation of materials and their applications
and Metal-Organic Frameworks have the potential to address these challenges, if the chemistry behind the materials can
be dominated. This report focuses on summarizing the many discoveries and rational approaches that chemists have taken
to improve the chemistry of Metal-Organic Frameworks, like modular chemistry, reticular synthesis, robust dynamics and
multivariable frameworks. Finally, a few examples of the applications of complex frameworks are briefly discussed to give
an idea of what the science has achieved, for the time I had to read the information available on the Internet.
Keywords: Metal-Organic Frameworks, modular chemistry, reticular synthesis, the materials beyond

Background: The origin of Metal-Organic Frameworks


For many years scientists have worked to design and develop molecular structures from chemical building blocks, as
if they were legos, because one can only imagine all the possible applications for chemical structures with subunits of
different chemical behaviors, like gas storage, energy storage, nanomaterials, catalysis and mimicking biochemical
reactions. This idea for a long time has had many limitations such as synthesis, chemical stability, thermal stability and
versatility. However, since the invention of metal organic frameworks (MOFs), there has been an increasing interest in the
potential uses and the rational synthesis and design of such materials.
One of the earliest examples of the idea that well characterized chemical subunits could be used to form molecular
frameworks dates from 1959, where bis(adiponitrile)copper(I) nitrate was synthesized. This complex consists of a Cu+
ion surrounded tetrahedrally by four nitrogen atoms, one coming from each adiponitrile molecule, organic subunits. The
unit cell of the compound can be described as follows: a copper atom at the origin of the cell is linked to four copper atoms
belonging of different unit cells1, (Fig. 1.A). In comparison to a diamond lattice structure, which is dense due to the link
between carbon atoms only, the one in bis(adiponitrile)copper(I) nitrate has longer distances between lattice points and
leaves enough space between them, making the crystal less dense than diamond; however, the structure could
interpenetrate itself, thus creating a variable porosity for the molecule. Some other examples include similar structures,
however, their porosity was dependent on “guest” molecules, as the ones of the solvent, that when diffused outside of the
empty space inside the framework molecule, the structure was altered. Later, Yoghi and his research group in UC Berkeley
in 1998 synthesized the molecule Zn(BDC)•(DMF)(H2O), where BDC is 1,4-Benzenedicarboxylate (terephthalate)
and DMF is N,N′-dimethylformamide, that was capable of conserving its molecular structure after solvent molecules left
the framework, and also could adsorb gas molecules to fill the empty space to give a Type-I isotherm that was typical for
porous materials like zeolites2. This was tested for N2 and CO2 and the framework could absorb in the range of 770-1500
mg/g. This new compound, in contrast to the previous framework, used multi-metal rigid joints (group of vertices) to
connect the organic linkers, instead of using just a single atom as a joint (one vertex) (Fig. 1.C), however the framework
in the figure uses Cu atoms instead of Zn atoms. Since then, it was proven that rigid MOFs could be synthesized, and they
could be used to store other materials inside, however their surface area was not as big as that found in zeolites or other
porous materials, so they were not as valuable.
The next new approach was to use these inorganic multi-metal joints and some organic linkers that could fit the lattice
of a cube, in order to explore the possibility of expanding the length of the linkers and increase the surface area of the
framework. The first attempt was MOF-5, a framework that used terephthalate as a ditopic linker, and zinc oxide Zn4O as
a multi-metal joint (Fig. 1.D). The lattice formed by this MOF had an exceptional surface area of 2900 m2/g (comparable
of the surface area of a soccer field), pores of about 1nm of diameter, and a rigid architecture. Also, it could adsorb amounts
of gases and liquids like Ar, CH2Cl2 and C6H6 in the range of 770-1500 mg per gram of MOF-5, at 77K. This was the
greatest surface area registered for a porous material at the time, in 19995, and it started the active research in MOFs.
After the context, a formal definition for MOFs may be introduced. MOFs are materials constructed by joining
inorganic metal containing units, also known as secondary building units (SBUs), with organic linkers, using strong bonds
made by reticular synthesis, to create open crystalline frameworks with permanent porosity6. In other words, these
materials are a type of three-dimensional rigid, sponge-like, symmetrical, chemical lattices, that have subunits made of
complexes between organic ditopic or polytopic linker molecules and joint molecules that contain a metal. A ditopic linker
is a minimal requirement to build MOFs because the framework only forms if the SBUs are joined in series, from two or
more sides, to form a symmetrical structure. The porosity of MOF crystals relies on strong bonds between linkers that
makes the crystal less dense, as opposed to other crystals that normally are compact because they are packed on top of one
another to form the crystal, which is denser. However, porosity is not the only feature that can be exploited in MOFs.
Nowadays, scientists focus on developing models and prototypes of MOFs that can have many complex functionalities by
adding new chemical reactivity to the pores by manipulation of the organic linkers, creating multivariable (MTV) MOFs,
and can be useful in areas such as computation and information processing and complex molecular dynamics, which most
likely resemble the behaviour of biomacromolecules like DNA, RNA and proteins. This report focuses on exploring the
new chemistry developed to achieve MTV-MOFs that perform highly complex dynamics through adding heterogeneity
to crystals and also on giving examples of this application in photochemical catalysis and nanomaterials.

A) C)

B) D)
Fig. 1. First insights into metal organic frameworks. A) Schematic diagram of a single 4-connected network in the crystal
of bis(adiponitrile)copper(I) nitrate with a distorted diamond lattice1. B) A diamond unit cell3. C) Synthesis and crystal
structure of Zn(BDC)•(DMF)(H2O)4. D) MOF-5 crystal for one cube and the two-dimensional grid of a series of
cubes on a plane5

Modular chemistry: nomenclature and reticular synthesis in MOFs


The development of the MOF chemistry is one of the fastest in the last century, since the appearance of MOF-5.
Many research groups have been expanding the porosity of MOFs, up to 7000 m2/g in 2012. Computationally, the
hypothetical theoretical limit for the surface area within a MOF was been calculated to be of around 11000 m2/g7. Also,
the industry has moved on and started the marketing of MOFs with promising outcomes. As an example, Rainer Zietlow,
a German adventurer, travelled for 6 months, a distance of 28000 miles across 40 countries, in the EcoFuel World Tour,
using a Volkswagen Caddy optimized to use natural gas as a fuel, stored in a MOF tank. This experiment proved that MOFs
are capable of resisting long distance reuse and exposure to moisture and impurities present in the natural gas. Also, there
are companies like BASF in Ludwigshafen, Germany that produces several types of MOFs in batches of 100 kg/day for
commercial and academic uses8. However, most of the MOFs synthesized today are used only in the laboratory scale due
to refining details of the materials that need to be addressed.

Nomenclature and definitions


In general terms, the chemistry of MOFs has advanced due to four major discoveries: the SBUs achieve their rigidness
from default geometric shapes or topologies, known as nets, like cubes and octahedra that can give a permanent porosity
to the MOFs, the rigidness of the SBUs can exploit the high porosity and high surface area of MOFs, the organic linkers
can be modified after synthesis to add other chemical species to change the chemical behaviour of each pore in a MOF,
and the post-synthesis modifications of MOFs can add variability to the framework in order to achieve complexity of
different levels in a predictable and rational manner6. The concept of bonding SBUs to organic linkers is analogous to
using discrete standard parts like exchangeable legos or modules (Fig. 2.A), which originate the concept of modular
chemistry. In terms of coordination chemistry, SBUs are rigid inorganic complexes and cluster entities that provide
coordination environments through many-vertices metal frameworks for organic ligands, that form in situ when the MOF
will be synthesized. Now, it is important to clarify that ditopic implies bidentate, however the opposite is not the case. In
contrast to chelating ligands, in MOFs, a ditopic molecule of organic ligand binds to two metal atoms, each from a different
side of the molecule. These ligands can still be multidentate, however they will not chelate the metal, unless one or more
of the polytopic sides in the molecule has more than one electron donating group or atom. For example, the In MOF called

3
InPF-13, with formula [In2(hfipbb)3(2,2’-bipy)2]•2H2O (H2hfipbb = 4,4’-hexafluoroisopropylidene bis benzoic acid;
bipy = bipyridine), has a ditopic linker with two chelating bidentate carboxylate groups, that act in a h2µ-h2µ mode9. The
process of replacing a one-atom vertex with a framework of many vertices (SBUs) in a net called decoration of the MOF.
These are called N-connected nets, where vertices are linked to N neighbours, and the length between the vertices is
constant and minimal, thus being called also N-coordinated spheres. The extended nets will therefore have many SBUs,
and when they are of the same type, the spheres are called uninodal. Nets in which all vertices, edges and angles are the
same are referred to as regular. If only edges are the same, then the net is referred to as quasirregular. However, optimal
porosity (non-interpenetrating structures) in a MOF can be obtained when all the SBUs have the same number of vertices
as the net itself. In other words, to have a N-connected net with SBUs of N-vertices. This is called an augmentation of the
MOF. To visualize this, one could imagine a cube, in which all vertices contain a Ca atom, coordinating identical linkers
between the vertices. The augmented cubic MOF would be one in which all vertices contain the CaB6 framework where
B6 units augment the 6-connected vertices, minimizing interpenetration. Once the porosity is optimal by geometry, it can
be increased by the process of expansion, which consists in increasing the length between the SBUs in the vertices of the
net, by using longer organic linkers, which could mean using many short organic linkers bonded in sequence or one longer
molecule11, for any polytopic linker used.

B)

A)

C) D)
Fig. 2. Versatility of MOFs. A) Inorganic SBUs and corresponding organic linkers (to the right) for augmented
framework6. B) Decorated and expanded framework12. C) General vertex figures in different kinds of nets13. D) SrSi2 net

Porosity and topology


Many extended combinations of MOFs can be achieved at near room temperature, if they can keep their structural
integrity throughout the process, the attraction mechanisms and the energy of the system. One can imagine the process of
making MOFs as triggering organometallic complexes to self-assemble into a framework. Self-assembly is about how
independent elements of a system come together on their own to build symmetrical, complex structures, that minimize
the energy when being together, into the most stable state possible. However, most of the approaches to synthesis have
been “shake and bake” chemistry rather than a rational approach to obtain desired materials. This lack of control during
the synthesis of solids occurs due to the structural changes that the precursors go through, and there is a poor geometrical
correlation between reactants and products, however, there are some clues. The complex Zn4O(BDC)3 can be taken as an
example. First, its SBUs consists of M-O-C (M: metal, O: oxygen, C: carbon) bonds that connect each other, then these
atoms define the points of extension of the framework, which will predict the overall topology of the MOF. Because every

4
O atom can complex to four Zn (II) atoms (Zn4O6+), and each metal per Zn4O6+ can bear three oxygen atoms coming
from the carboxylate molecules (BDC), being tetrahedral around the metal, giving a total of 6 BDC (2 oxygen atoms each)
that will complex to Zn4O6+, which is octahedral around the SBU (Fig. 2.B). Second, the choice of ligands is important
because it can help predict the shape of the pores and also the geometrical disposition of the SBUs in space after
polymerization12. Third, solution synthesis and crystallization of modular solids can be achieved by in situ formation of the
SBU, which requires a copolymerization between a transition metal ion and a polytopic organic linker. Generally, this takes
place in a solution with the acid form of the linker and an appropriate metal salt, where a base, like an amine is added slowly
to deprotonate de acid and start the assembly of the MOF. The nucleation rates and number of nucleation sites can be
controlled with diffusion of base, solvent polarity and concentration gradients. This results in the formation of linked
cluster entities, as the case of the complex Zn4O(BDC)3, by Yaghi in 1998, which used Zn(II) and ditopic terephthalate
anions as precursors, that led to the formation of the complex, with SBUs of the formula Zn2(–COO)4(H2O)2, which is a
highly symmetrical structure of the Td point group. The two-dimensional grid for Zn4O(BDC)3, (Fig. 2.B), is held together
by strong hydrogen bonding interactions among water molecules and carboxylate linkers12, 13.
There are several synthetic, rationally designed, framework topologies that authors like Yaghi and Eddaoudi have
characterized, by single-crystal X-ray diffraction data, to be the most frequent in MOFs. They have found that it is expected
to form highly symmetrical structures, with rather simple coordinating figures and coordination numbers, due to the lack
of control of the synthetic mechanisms (Fig. 2.C). For example, there are only 52 known 3-connected nets, of which only
one has a 3-fold axis, with all angles equivalent, thus, being the only regular net. That is the case of SrSi2 (Fig. 2.D) and
other frameworks with the same topology. Their vertices have a D3 point symmetry, and are chiral. Zn2(BTC)NO3
framework adopts this structure, in which Zn(–CO2)3 SBUs and benzene linkers decorate the vertices to yield a symmetry
equal to the Si net of SrSi2. In the case of 4-connected nets, each vertex must have a D2 symmetry or higher to be uninodal.
To be regular (where all angles are related by symmetry) the vertices must have a Td symmetry, such as the SBU in
Zn4O(BDC)3, and such as diamond (Fig. 1.B). There are other types of 4-connected nets, such as the complicated sodalite,
or the one for NbO, which could be imagined as three squares, each one on a plane that lies on each axis of a three
coordinate Cartesian system (x,y,z), where each vertex of each square and each intersection between squares contains an
atom. There are many more complex structures to describe, but fall out of the scope of this work. In general, rational
synthesis of these molecules will mostly produce simple symmetrical structures and one might only stumble upon much
more complex structures, by chance only13. Now, one can take these “standard” simple geometries to mimic the structure
of a previously characterized net, in order to decorate a framework, in other words, to change its vertices by any SBU. On
reported approach was to take a paddle-wheel cluster Cu(CO2)4 (the SBU of the MOF showed in Fig. 1.B), derived from
Cu(II) carboxylate, that has a square coordination figure, with adamandane cluster as a second SBU, that has a tetrahedron
coordination figure, to synthesize MOF-11. The combination of a square and a tetrahedron should, in principle, give the
crystal structure of cooperate (PtS) (Fig. 2.C and Fig. 3.A). Indeed, that was the obtained structure for the mixture. Other
interesting property of the crystal is that it has open metal (OM) sites when the solvent comes off the framework, leaving
it anhydrous. The magnetic experiments of the framework show that the Cu(II) carboxylate has a greater
antiferromagnetic coupling (2J = –444 cm-1) inside the framework than it would have as an isolated species (2J = –280 cm-
1
). After full characterization, MOF-11 showed good potential to be used as a sensor, because if the metal was replaced by
an actinide, the complex could be luminescent in the presence of small molecules, like water and ammonia. Other
applications are in catalysis and to have functionalized linkers after modification. However, the best example there is to
start with for an optimal and rational design of a framework is MOF-5, as explained before. Finally, the last
recommendation that reports give is that to always maximize the porosity of the frameworks, interpenetration must be
minimized as explained before, but, according to a model of the volume of SBUs approximated to a sphere, four factors
must be taken into account: the volume percentage occupied by interpenetrating structures, the diameter d of the spherical
SBU, the length l of the linkers and the number n of frameworks that interpenetrate (Fig. 3.B). The interpenetration n is
directly correlated to l, and the free volume inside the framework is related to d. Then, to maximize porosity, the
recommendation is to always use relatively larger SBUs and keeping the diameter of the pores small, bellow 2 nm, by
keeping the length of the MOFs low, which can be achieved by using polytopic linkers without losing distance between
the SBUs.

Reticular synthesis
It is important to stablish that SBUs are not isolatable entities and that is why they are synthesized in situ, however
organic linkers are, so using molecules as legos is not a real possibility; however, the rational, logical approach is as
described before, with MOF-11. The designer first synthesises conceptually MOFs by choosing a specific target network,
which is then deconstructed into its geometric units. After that, molecular building blocks (SBUs and linkers) with the
same geometry as the network topology are chosen to assemble into a MOF structure. As an example, if the designer would
want to synthesise a network with the diamond topology, the smallest and simplest subunit of the yields tetrahedral
building blocks. Thus, SBUs and linkers with tetrahedral building blocks must be chosen. However, most crystals will not
be as simple to deconstruct as diamond and may require a more complex analysis of the possible subunits that would yield

5
the desired crystal. This is the case of lonsdaleite, the “hexagonal diamond”, that has a similar crystal structure as diamond
and could use tetrahedral subunits, but diamond will be the standard crystal for tetrahedral subunits because it is more
symmetrical than lonsdaleite. If its crystal was analysed, it could have boat like and chair like subunits to form the crystal
(Fig. 3.C.). Other examples include MOP-1 and MOF-222, which use a paddle wheel with one particular linker to achieve
complex crystal three-dimensional and two-dimensional geometries (Fig. 3.D.). This explanation pretty much sums up
the reticular synthesis principle14. It is called reticular because it is based on a pre-known network as the starting point.
Perhaps, it would have been a really hard task to know which subunits to use to form a desired MOF, because one would
have to do crystallography experiments and characterizations before knowing the possible structures of a desired crystal,
however, today there are thousands of subunits characterized in Cambridge Structure Database (CSD) that help to
computationally predict the most optimal subunits to form MOFs.

A) B)

C) D)
Fig. 3. Considerations in reticular synthesis of MOFs. A) Crystal structure formed by Cu(CO2)4 and adamandane as of
cooperate (PtS)13. B) Mathematical model of the percentage volume occupied (lighter zones represent greater
percentage) by interpenetrating frameworks as a function of d, l and n13. C) Geometry of lonsdaleite14. D) MOP-1 and
MOF-222 used reticular synthesis to be created14.

The Materials Beyond: Complex arrays for complex applications


The literature of MOFs has already exploded and reached a maturity that lets scientists think of new ways to improve
the materials and add complexity to their functionalities. Most of the applications are pretty known to humans, as the
Volkswagen Caddy MOF tank example, but some other new could be MOFs having compartments linked together to
operate separately and synergistically, carry out parallel operations, ability to count, sort, and code information, and
capability of precise molecular dynamics. One readily available example of molecular machines known to man are
biomacromolecules like proteins, DNA and RNA, which indeed code information, carry out complex dynamics and can
compute complex algorithms like inducible modular gene expression, a research topic in the emerging field of synthetic
biology6,15. The most important contrast between MOF crystals and these molecular machines is their self-similarity.
Crystals are rather symmetrical and have subunits that repeat themselves. This characteristic prevents MOFs from
achieving complex tasks as those mentioned, but is still of great value due to the stability it provides. Therefore, there is the
need to figure out ways to add heterogeneity to the frameworks in order to expand their functionalities. This is rather a
challenge for reticular synthesis because the most likely outcome of mixing many subunits and placing them together by
synthesis is a mixed phases network, rather than a single phase with mixed components, and the exceptional properties of
MOFs may be lost, because they are flexible. However, if the chemistry can be tricked in order to build these materials,
there are unlimited possibilities for their use, and these are called the materials beyond6. There have been attempts to deal
with the limitations, and are explained bellow.

6
Robust dynamics
Dynamics in a molecule are of importance in applications like catalysis and mechanical operations like torsion and
bending, as the ones of kinesin, the walking protein. Some chemists have idealized the way in which dynamics can be added
to a rigid MOF structure in order to enhance its functionalities. The idea sums up to covalently insert into the framework
linkers some mechanically interlocked molecules that are capable of dynamics, maintaining the MOF’s robustness, in other
words, the integrity of the framework. The molecules that are capable dynamics are therefore capable of random motion
(Fig. 4.A), and must be trapped somehow, inside a three-dimensional structure, to perform their functions without
compromising the MOF16. There have been advances in the field of molecular machines, whose components were held
together mechanically, by clipping or other mechanism; however, up to 2010, these molecules described motion in an
incoherent manner in solution or in condensed phases. Maybe the rigid backbones provided by frameworks could be an
interesting solution to compartmentalize the motion of the molecular machines. There is where the name “robust
dynamics” comes from. Molecules like rotaxanes and catenanes capable of dynamics have been characterized as molecules
added to surfaces by self-assembly methods, in order to observe their performance, but it was found that the dynamics of
these machines dropped dramatically due to loss of function. MOFs, on the other side, have been good platforms for the
dynamics of rotaxanes, achieving pseudorotaxanes, that contain a 36-membered polyether ring around a linker (MOF-
1002), a sort of receptor for the ring to covalently attach it to the linker of the MOF. In MOF-1002 the ring is disordered
and moves around space in a poorly controlled manner, due to planar chirality (Fig. 4.A). The disorder is not desired
because it lowers the capability of controlling the dynamics inside the MOF cells. This chirality occurs because there is no
control over the enantiomeric forms during their synthesis of MOF-1002. One possible strategy to solve the disorder is to
use asymmetric catalysis, to produce homochiral MOFs. Another form of controlling the dynamics in a MOF is to use
mechanically interlocked components, as pure rotaxanes or catenanes have, or as biomacromolecules have, like DNA
Polymerase III, which operates by placing its polymeric ring around the DNA molecule as a thoroidal clamp with the
catalytic subunit on the DNA backbone. There is an example of a rotaxane macrocycle containing the Mn(III) porphyrin
that catalyzed epoxidation of a polybutadiene substrate as a DNA Polymerase or other DNA binding protein would modify
the DNA. This type of functionality has not been able to add into a MOF, however things like molecular shuttles have been
incorporated successfully into MOFs17 (Fig. 4.B), showing the possibility of adding complexity with dynamics.

Defects, disorder and flexibility


In terms of the physical properties of MOFs, they are different than other coordination compounds, because these
materials are less flexible than expected. Normally, coordination polymers have dynamic behaviors, however there is a
macroscopic rigidness to MOFs, like negative thermal expansion (contraction under the exposure to heat), auxeticity (the
material becomes thicker when stretched along certain directions) and negative linear compressibility (total volume
reduction with expansion along a specific direction under mechanical pressure). These manifestations could be explained
by the wine-rack, honeycomb and related topologies of MOFs. There are some defects associated to MOFs also, like the
incomplete activation of its porosity, which can be observed when computational and experimental predictions have
different ranges in the porosity measured (or calculated). There are other defects like metal reduction or misplacement
and node or linker vacancies. Even when these defects can be seen as bad for the framework performance, they can
potentially increase functionality if they can be engineered. Actually, having MOFs with defects is a standard, due to the
many errors that can occur during synthesis. These have created rather disordered (to some degree) MOFs18.
Defects, disorder and flexibility occurrence in a MOF can be linked to entropy (Fig. 4.C). For example, the
[NH4][Zn(HCOO)3] framework is a ferroelectric (spontaneous electric polarization, reversible under an external electric
field, due to permanent dipole moment inside the crystal) structure at low temperatures and adopts the polar hexagonal
space group P63. However, when heated to 192 K, it undergoes a phase change, acquires the nonpolar hexagonal space
group P6322 (more disordered), becomes paraelectric (no permanent electric dipole) due to the dynamical disorder of
NH4+ ions inside the framework, changes in hydrogen bonding occur between the order-disorder transition, and
undergoes an elastic softening. These properties have been studied with resonant ultrasound spectroscopy in the
temperature range between 10 K and 300 K18, 19. Another case of study is the UiO-66 family of the uninodal frameworks,
with the formula [M6O4(OH)4][C6H4 (COO)2]6, where M is a metal, in this case Zr coordinated by 12 organic linkers,
and which can have up to 10% of terephthalate linkers missing. This defect (or deflection from perfect symmetry) leads to
an increase in surface area from 1000 to 1600 m2 g–1, a reduction in the framework connection from 12 to almost 11, leaving
metals coordinately unsaturated, thus enhancing Lewis acid catalyzed reactions in the SBUs and adsorption capacity of
the whole framework, and also allows he presence of hydroxylated linkers which are non-air sensitive. Also, properties like
auxeticity and anisotropy were found to increase when the number of likers was gradually reduced 20. Now two different
likers (NO2 and MeO) were mixed in with when one kind of SBU to form the framework of the formula {Zn(5-linker-
ip)(bpy)}(0.5DMF·0.5MeOH), (ip = isophthalate, bpy = 4,4ʹ-bipyridyl), a flexible pore opening that changed from open
to closed was created, increasing the selectivity for CO2 gas adsorption (from a mixture of CO2/CH4) when compared to
only using one kind of linker21.

7
A) B)

C)

D) E)
Fig. 4. Frontiers and achievement for the materials beyond. A) Pseudorotaxanes in random motion integrated in a
MOF16 B) Shuttle molecular machine integrated in a MOF17. C) The role of entropy in defects, disorder and flexibility
of frameworks18. D) [Zn(Gly-Ala)2]•(solvent) dipeptide MOF24 E) MTV-MOF-5 scheme22

8
Defects and disorder are linked when the nucleation of frameworks is examined. For example, [Al(OH)(BDC)]
framework starts a slow initial self-assembly into nanoparticles of around 10 nm of diameter, and then proceeds rapidly to
form extended structures up to 200 nm particles. If the process of the larger particles is disrupted, by the addition of other
linkers, metals, or heat to the system, then defects are created and disorder within the framework appear in the form of a
gel, generally with the same structure of the non-disrupted framework, but with changed physical properties, due to these
“impurities”, like lower melting points18. However, the next logical step is to try to trick these defects to improve properties
and enhance functionality of the frameworks, rather than trying to eradicate them.

Multivariate (MTV) MOFs


In general, the new approach to MOF design is to add functional groups in any possible form in order to achieve
better, more complex materials, which are referred to as multivariate MOFs. The functional groups in these designs have
a particular, preconceived orientation, number, relative position, and ratio along the backbone, and point to the centre of
the pores in the MOF in order to spatially access any molecules diffusing inside of to minimize the interpenetration inside
the cavities of the MOF, according to the rules of reticular synthesis (Fig. 4.E). The most famous example of rationally
designed MTV-MOFs was MTV-MOF-5-ABCD (A, 1,4-benzenedicarboxylate; B, -NH2; C, -Br; D, -(Cl)2; E, -NO2). It
was made by adding 1,4-benzenedicarboxylate to a liquid mixture of desired groups and then polymerized adding a base,
as noramlly. Surprisingly, the structure of the MOF was highly crystalline, heat resistant (no weight loss up to 400oC), of
rigid porosity, and a composition of functional groups of 1.00: 0.12: 0.56: 0.40 (A: B: C: D, respectively) was observed
with data from powder x-ray diffraction, 1H NMR on acid-digested solutions and thermogravimetric analysis. 13C
CP/MAS NMR spectra of empty MTV-MOF-5-ABCD showed resonances at 150.3, 127.0, 133.7, and 136.3 parts per
million (ppm), which correspond to the carbon atoms of each functional group22. Each functional group was also analyzed
in separate studies and it was found that the same NMR results were obtained, with a shift of only 2 ppm, indicating that
there was no unbound organic link within the network. Now, the mapping of the functional groups within a MOF is hard
to determine with precision with X-ray diffraction, but can be explored with solid-state nuclear magnetic resonance
(SSNMR) 13C NMR for relative pairwise interactions between elements throughout the whole sample, rotational echo
double-resonance (REDOR) 13C NMR for quantitative distances between atoms of the crystals, in combination with
computer simulations, to achieve a great three-dimensional approximation of the distribution of the functional groups.
The idea consists of linking Monte Carlo and molecular dynamics simulations to the NMR data for error estimations and
parameter fitting. This is important because it is thought that reticular synthesis approaches to MOF design and
production may deviate from the desired uniformity and periodicity of the functional groups, like random, alternating or
various cluster formations. It was found that all these structures could be achieved, depending on the linker used and the
substituents in it (Fig. 5.A)24. Therefore, with these new evidence and tools, the research of MOFs could become more
meticulous in order to control and monitor the disorder, defects and variability during synthesis.

B)

A) C)
Fig. 5. Variability in MOFs. A) Calculated and experimental REDOR NMR results for MTV-MOF-523. B) Confinement
of Photosystem I in ZIF-8 MOF. C) Crystals of PSI-ZIF-8 complex 3.

9
A really interesting example of molecular dynamics in a MOF was reported in 2010, where dipeptide linkers were
used to build a MOF with variable porosity, that resembled the dynamics of a protein undergoing conformation change.
Experiments aided by molecular dynamics simulations showed that low-energy torsions and displacements of the peptides
enabled the available pore volume that increase in response to the load of the guest molecules that diffused inside the pores.
Proteins are characterized by an adaptable conformational change in response to their environmental stimuli. This inspired
the adaptation of MOFs to form large ensemble of energetically low-lying and kinetically accessible states. The study
obtained crystals of the metal-dipeptide framework [Zn(Gly-Ala)2]•(solvent), were obtained by reacting zinc nitrate with
Gly-Ala in 90% methanol. The zinc ions are tetrahedrally coordinated (fig. S6A and table S4) to four dipeptide ligands.
ligands; two dipeptide ligands are coordinated by the C-terminal Ala carboxylate groups and two by N-terminal Gly amine
groups. The amide groups line the walls of the pore and are linked by hydrogen bonds with an alignment similar to that in
b-sheet structures of polypeptides, controlled by the torsional conformation of the dipeptide, which is defined by four
torsion angles (Fig. 4.D). The ability of these angles to rotate and change plays an important role in the variable porosity
of the MOF. If one of these was fixed, steric interactions between methyl groups on opposite sides of the pores would
prevent the pore from closing23.
Another example is the use of a MOF to confinement of Photosystem I (PSI), a transmembrane protein in the
tilacoids of leaves in plants, that is able to perform photoactivated charge separation with 1 V driving potential and 100%
quantum efficiency during the photosynthetic process, which makes it a potential candidate for integration into bio-hybrid
solar energy harvesting devices. However, the enzyme alone cannot be efficiently used, unless it is integrated into a
framework that supports it physically. MOFs are the kind of network three-dimensional material that can be optically
active and could help overcome the challenge. That is exactly what was done. A study investigated the optical response
and photoactive properties of PSI encapsulated in a ZIF-8 MOF, because it could provide a scaffold with a robust structure
to protect the enzyme’s delicate machinery, it’s precisely coordinated chlorophyll networks, from denaturing agents. The
protein was obtained Thermosynechococcus elongatus. Then ZIF-8 precursor Zinc acetate dehydrate was mixed with the
enzyme and then the linker 2-methylimidazole (Hmim) was added, also in the presence of the polymerizing agent. This
MOF with the confined enzyme was used directly in experiments. Nucleation of ZIF-8 creates a dense solution of highly
concentrated ZIF-8 nanoparticles that scatter light, thereby blocking the light and preventing photoactivity measurements.
To solve this, concentrations were standardized to promote PSI surfaces to be the primary nucleation sites, conserving
high crystallinity. Also, when the optical measurements were made, there were differences. Typically, PSI has two
characteristic absorption peaks at 440 nm and 680 nm. Blue shifts in the fluorescence emissions from UV-vis (680 nm to
676 nm), due to the ZIF-8 scattering, and proved the successful confinement of PSI in ZIF-8. Pump–probe spectroscopy
was used to measure the changes in the optical constants as a function of time delay between the arrival of a pump detector
and the probe pulses. This information describes the relaxation of electronic states in the enzyme. The measurements
confirmed the photo activity of the PSI-ZIF-8 complex confirmed by an artificial external charge transfer of P700+ complex
and an electron acceptor. The system shows that photo activated cages in MOFs can actually be used for applications like
bio-hybrid solar energy harvesting devices.

Conclusion
MOFs are molecules that have a great potential for uncountable applications, mostly because most of the
applications have not been explored or thought of yet. In the beginning, I thought these molecules were much simpler than
they really are, and also thought that their applications were not so many. However, this research has revealed valuable
information that helped me understand the importance of coordination chemistry and the really exiting field of inorganic
chemistry. MOFs are only one small part of all the science behind materials and it is exiting. I could only imagine how a
country like Ecuador could benefit from having infrastructure, human talent and resources to exploit materials sciences,
and specifically the use of MOFs in industry to generate value to chemical industrial processes and potentially lower costs
of production and promote technological development. Specifically referring to MOFs, these molecules need to be studied
in order to exploit their capabilities, mostly because there are already many ideas in the field that can help develop potential
applications for MOFs, but also because these materials could be tuned to have astonishing functionalities. There are many
problems us humans should address with the scientific method in order to reduce detrimental consequences of our activity,
like global warming, garbage management, poverty and many more. The versatility of MOFs could, perhaps, contribute to
use resources in an intelligent manner.

10
References
(1) Kinoshita, Y.; Matsubara, I.; Higuchi, T.; Saito, Y. Bulletin of the (14) Yaghi, O. M., O'keeffe, M., Ockwig, N. W., Chae, H. K., Eddaoudi,
Chemical Society of Japan1959, 32(11), 1221–1226. doi: M., & Kim, J. Reticular synthesis and the design of new
https://doi.org/10.1246/bcsj.32.1221 materials. Nature2003, 423(6941), 705.
(2) Li, H.; Eddaoudi, M.; Groy, T. L.; Yaghi, O. M. Journal of the (15) Cheng, A. A., & Lu, T. K. Synthetic biology: an emerging
American Chemical Society1998, 120(33), 8571–8572. doi: engineering discipline. Annual review of biomedical
10.1021/ja981669x engineering2012, 14, 155-178.
(3) University of Bristol. Diamond. (16) Deng, H., Olson, M. A., Stoddart, J. F., & Yaghi, O. M. Robust
http://www.chm.bris.ac.uk/motm/diamond/diamondh.htm dynamics. Nature chemistry2010, 2(6), 439.
(accessed Dec 7, 2018). (17) Lewis, J. E., Galli, M., & Goldup, S. M. Properties and emerging
(4) Yaghi, O. https://youtu.be/bzDIH7olP0c Molecular Foundry, applications of mechanically interlocked ligands. Chemical
Berkeley Lab. (accessed Dec 4, 2018). Communications2017, 53(2), 298-312.
(5) Li, H.; Eddaoudi, M.; Okeeffe, M.; Yaghi, O. (18) Bennett, T. D., Cheetham, A. K., Fuchs, A. H., & Coudert, F. X.
M. Nature1999, 402(6759), 276–279. doi: Interplay between defects, disorder and flexibility in metal-organic
https://doi.org/10.1038/46248 frameworks. Nature chemistry2017, 9(1), 11.
(6) Furukawa, H.; Cordova, K. E.; Okeeffe, M.; Yaghi, O. (19) Zhang, Z., Li, W., Carpenter, M. A., Howard, C. J., & Cheetham, A.
M. Science2013, 341(6149), 1230444–1230444. doi: K. Elastic properties and acoustic dissipation associated with a
10.1126/science.1230444 disorder–order ferroelectric transition in a metal–organic
(7) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, C. E.; framework. CrystEngComm2015, 17(2), 370-374.
Sarjeant, A. A.; Snurr, R. Q.; Nguyen, S. T.; Yazaydın, A. Ö.; Hupp, (20) Wu, H., Chua, Y. S., Krungleviciute, V., Tyagi, M., Chen, P.,
J. T. Journal of the American Chemical Society2012, 134(36), Yildirim, T., & Zhou, W. Unusual and highly tunable missing-linker
15016–15021. doi: 10.1021/ja3055639 defects in zirconium metal–organic framework UiO-66 and their
(8) Jacoby, M. Chemical & Engineering News2008, 86(34), 13–16. important effects on gas adsorption. Journal of the American
doi: 10.1021/cen-v086n034.p013 Chemical Society2013, 135(28), 10525-10532.
(9) Aguirre-Díaz, L. M.; Iglesias, M.; Snejko, N.; Gutiérrez-Puebla, E.; (21) Cohen, S. M. Postsynthetic methods for the functionalization of
Monge, M. Á. RSC Advances2015, 5(10), 7058–7065. doi: metal–organic frameworks. Chemical reviews2011, 112(2), 970-
10.1039/c4ra13924k 1000.
(10) Aguirre-Díaz, L. M., Iglesias, M., Snejko, N., Gutiérrez-Puebla, E., (22) Helal, A., Yamani, Z. H., Cordova, K. E., & Yaghi, O. M.
& Monge, M. A. Toward understanding the structure–catalyst Multivariate metal-organic frameworks. National Science
activity relationship of new indium MOFs as catalysts for solvent- Review2017, 4(3), 296-298.
free ketone cyanosilylation. RSC Advances2015, 5(10), 7058- (23) Rabone, J., Yue, Y. F., Chong, S. Y., Stylianou, K. C., Bacsa, J.,
7065. Bradshaw, D., ... & Wiper, P. (2010). An adaptable peptide-based
(11) O'Keeffe, M., Eddaoudi, M., Li, H., Reineke, T., & Yaghi, O. M. porous material. Science, 329(5995), 1053-1057.
Frameworks for extended solids: geometrical design (24) Kong, X., Deng, H., Yan, F., Kim, J., Swisher, J. A., Smit, B., ... &
principles. Jour. Sol. S. Chem.2000, 152(1), 3-20. Reimer, J. A. (2013). Mapping of functional groups in metal-
(12) Eddaoudi, M., Moler, D. B., Li, H., Chen, B., Reineke, T. M., organic frameworks. Science, 341(6148), 882-885.
O'keeffe, M., & Yaghi, O. M. Modular chemistry: secondary (25) Bennett, T. H., Vaughn, M. D., Davari, S. A., Park, K.,
building units as a basis for the design of highly porous and robust Mukherjee, D., & Khomami, B. (2019). Jolly green MOF:
metal− organic carboxylate frameworks. Acc. Chem.
confinement and photoactivation of photosystem I in a
Res.2001, 34(4), 319-330.
(13) Schoedel, A., & Yaghi, O. M. Porosity in Metal–Organic
metal–organic framework. Nanoscale Advances.
Compounds. Macrocyclic and Supramolecular Chemistry: How
Izatt-Christensen Award Winners Shaped the Field2016, 200

11

Anda mungkin juga menyukai