Next Article in Journal
Upper Critical Solution Temperature (UCST) Behavior of Coacervate of Cationic Protamine and Multivalent Anions
Next Article in Special Issue
Polyvinylnorbornene Gas Separation Membranes
Previous Article in Journal
Recent Progress of Carbon Dot Precursors and Photocatalysis Applications
Previous Article in Special Issue
Synthesis and Gas-Permeation Characterization of a Novel High-Surface Area Polyamide Derived from 1,3,6,8-Tetramethyl-2,7-diaminotriptycene: Towards Polyamides of Intrinsic Microporosity (PIM-PAs)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Highly Porous Organic Polymers for Hydrogen Fuel Storage

Department of Chemistry and Biochemistry, California State University, San Bernardino, CA 5500, USA
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(4), 690; https://doi.org/10.3390/polym11040690
Submission received: 1 February 2019 / Revised: 24 March 2019 / Accepted: 28 March 2019 / Published: 16 April 2019

Abstract

:
Hydrogen (H2) is one of the best candidates to replace current petroleum energy resources due to its rich abundance and clean combustion. However, the storage of H2 presents a major challenge. There are two methods for storing H2 fuel, chemical and physical, both of which have some advantages and disadvantages. In physical storage, highly porous organic polymers are of particular interest, since they are low cost, easy to scale up, metal-free, and environmentally friendly. In this review, highly porous polymers for H2 fuel storage are examined from five perspectives: (a) brief comparison of H2 storage in highly porous polymers and other storage media; (b) theoretical considerations of the physical storage of H2 molecules in porous polymers; (c) H2 storage in different classes of highly porous organic polymers; (d) characterization of microporosity in these polymers; and (e) future developments for highly porous organic polymers for H2 fuel storage. These topics will provide an introductory overview of highly porous organic polymers in H2 fuel storage.

Graphical Abstract

1. Introduction

The world is reaching its “Oil Peak” and petroleum resources are expected to be exhausted in coming decades. Seeking alternatives to fossil fuels is imperative. H2 is thought to be a viable alternative fuel due to its large abundance and clean combustion. In addition, H2 has a much higher energy density (142 kJ/g) than that of petroleum oil (47 kJ/g) [1]. However, the use of H2 as a fuel in automobiles has a major obstacle: the onboard storage of H2 gas. Many methods to store hydrogen have been proposed. The most direct way is to storage H2 as a liquid or a high-pressure gas [2,3]. This approach requires significant energy for liquefying or pressurizing H2 gas, which has a very low boiling temperature (20 K) and critical temperature (33 K), and also poses many safety concerns due to extremely low temperature and high pressure. In addition, H2 can embrittle steel gas tanks after long storage, generating additional risk. As an alternative, H2 storage media have been extensively explored, under the H2 program initiated by the U.S. Department of Energy (DOE) in 2003. The DOE set several criteria for materials for onboard H2 storage: (1) high storage capability, that is, 5.5 wt % and 40 g/L at ambient conditions; (2) rapid H2 release and recharge under moderate conditions; and (3) long recycling life, that is, more than 1000 recharge and discharge cycles [4,5]. The first two criteria have proved particular challenging, so that the DOE has postponed its target date several times. For instance, the DOE set its initial target for 2015, extended it to 2017, and then to 2020 [6]. The ultimate goal of the DOE is 6.5 wt % and 60 g/L by 2050, as summarized in Table 1 [7].
As seen from Table 1, there are many engineering and economic requirements, as well as scientific requirements for hydrogen fuel to be successfully employed as an onboard energy power for vehicles. In order to be adopted commercially, a hydrogen storage medium must meet all above criteria. Nevertheless, two fundamental scientific criteria, gravimetric capacity, and volumetric capacity, are the primary and most important parameters for scientists to pursue. Therefore, in this review, we will focus on progress made in meeting the first two criteria, particularly the gravimetric capacity, using porous organic polymers for the hydrogen storage.
Currently there are two major ways to store H2 molecules: (a) chemical absorption by forming hydrogen-containing molecules and (b) physical adsorption in highly porous materials. The former approach includes metal hydrides such as NaH, LiH, NaAlH4 [8,9,10], etc. as well as other hydrogen-containing molecules such as H3N–BH3 [11]. These materials store H2 via chemical bonding. The advantage of these materials is that they have a relatively high storage capability. The disadvantage is that they require high temperatures to break the chemical bonds and release H2 molecules, so that the recovery of H2 gas from these chemicals is not energy efficient. Organic hydrogen carriers or organic hydrides are quite reactive, leading to safety concerns, and disposal of metals poses the environmental concerns. In contrast, physical adsorption uses weak van der Waals interactions, and therefore H2 molecules can be released easily under moderate conditions. However, due to these weak interactions, physical adsorption, which can retain relatively large amount of H2 at liquid nitrogen temperature, has greatly reduced storage capability at room temperature. Physical adsorption methods can be achieved with materials with high porosity, including metal organic frameworks (MOFs), covalent organic frameworks (COFs), activated carbons, carbon nanotubes, carbides, and highly porous polymers [12,13,14]. Currently, the general approach for using physical adsorption in H2 storage is to increase the internal surface area in materials. For instance, activated carbons can have surface areas of up to 2000 m2/g and store H2 up to 5 wt % at 77 K/40 bar [15], but storage capacity decreases to 1.0 wt % at 298 K/200 bar [16]. Single walled and multiwalled nanotubes can store 6 wt % at 77 K [17]. MOFs, which are crystalline solids consisting of multidentate organic ligands connecting to a metal ion, afford a very large internal surface area, e.g., 5000 m2/g [18]. A wide range of MOF materials have been synthesized.
Porous organic polymers have great advantages over than other materials; for instance, highly porous polymers are more stable at the ambient conditions under which the metal ions in MOFs are sensitive the moisture. The polymeric structures can be well controlled via different organic synthetic routes and starting materials. Four groups of highly porous polymers have been studied intensively for hydrogen storage: (a) hypercrosslinked polymers (HCPs), (b) polymers of intrinsic microporosity (PIMs), (c) conjugated microporous polymers (CMPs), and (d) porous aromatic frameworks (PAFs). There are many excellent studies in this field that are making gradual progress towards the DOE hydrogen storage targets. Broom et al. provide an overview on porous materials for hydrogen storage, including practical considerations of the technology [19]. This review is not a summary of all of these works, rather, it starts from basic theory for H2 storage, and an overview of each storage medium, highlighting the most effective results, and then addresses the challenges and directions for developing highly porous polymers needed to meet DOE standards. The paper is arranged in the following structure: first, basic theoretical considerations for storing the maximum amount of H2 at ambient conditions are introduced; then the different types of highly porous polymers are summarized in terms their structure, surface area, porosity characterization, and hydrogen storage ability. Finally, future directions for investigation and possible effective solutions are proposed.

2. Theoretical Considerations

Theoretical background is introduced from two perspectives: first, we present the fundamental thermodynamic requirements for adsorption enthalpy needed to store the maximum amount of hydrogen, as well as releasing/adsorbing H2 molecules readily at ambient conditions as required by the DOE criteria; and second, we outline the enhancement of interaction between H2 molecules and the host cavity, using molecular energy level considerations, necessary to achieve the overall adsorption enthalpy required.
Probably the most promising theoretical work on the physisorption of H2 from the thermodynamic perspective was carried out by Bhatia et al., who were motivated by the conflicting reports on the capacity of H2 storage in highly porous materials [20]. For instance, an isoreticular metal organic framework (IRMOF) material, IRMOF-8, was first reported to have an H2 absorption capacity of 2.0 wt % H2 at 298 K/10 bar [21]. However, subsequent computer simulations from Rowsell et al. predicted a H2 uptake of only 0.75 wt % at 77 K/1 bar for IRMOF-8, suggesting a significantly lower capacity [22]. In order to explore the theoretical limit for the maximum delivery of H2, which is the difference between H2 in a charged and discharged material, Bhatia et al. used the Langmuir model to model H2 adsorption inside macropores/mesopores (pores with diameters of 2–50 nm and of >50 nm, respectively) of materials to find the amount of H2 adsorbed under equilibrium conditions, with an equilibrium constant K as shown in Equation (1).
n = K P n m 1 + K P ,
where K is equilibrium constant of the adsorption, P is the pressure of H2 gases, n is the amount of H2 adsorbed and nm is the quantity of active adsorbing sites in a material.
So the delivery (D) of H2 under two different pressures, P1 and P2, is:
D ( K , P 1 , P 2 ) = n 1 n 2 = K P 1 n m 1 + K P 1 K P 2 n m 1 + K P 2 .
When P1 and P2 are fixed, the equilibrium constant K for the maximum delivery (D) can be found by setting dD/dK = 0 in Equation (2) to give Equation (3):
K = 1 P 1 P 2 .
According to the Gibbs equation:
lnK = −ΔG°/RT = − (ΔH°− TΔS°)/RT.
By combining Equations (3) and (4), the optimal adsorption Δ H a d for the maximum delivery (D) is obtained as Equation (5):
Δ H o p t = T Δ S = R T 2 ln ( P 1 P 2 P 0 2 ) .
Equation (5) can be used to search for the optimal Δ H a d for different adsorption materials. On the other hand, for the same material, the optimal adsorption temperature Δ T o p t can be determined by rearranging Equation (5) to give Equation (6):
T o p t = Δ H [ Δ S + ( R / 2 ) ln ( P 1 P 2 / P o 2 ) ] ,
where ΔS° ≈ −8R for H2 adsorption in Langmuir model. If H2 is charged/discharged at the pressures of 30/1.5 bar and the temperature of 77 K, Δ H o p t is calculated to be –6.3 kJ/mol, close to –5.0 kJ/mol reported recently for a variety of porous materials at cryogenic temperature [23]. The difference between these values might come from the recharge P1 and discharge pressure P2. However, when storing H2 under the same pressures, but at 298 K as DOE expects, the calculated Δ H o p t is –15.1 kJ/mol. This value for the adsorption is optimal with respect to the affinity of hydrogen—strong enough to store a large amount of hydrogen gas at the charging pressure (~30 bar) but weak enough to release most of that hydrogen at the discharge pressure (~1.5 bar). Unfortunately, for most physical adsorbents, the heat of adsorption is much less that this value. For instance, for activated carbons or hydrocarbons, the adsorption enthalpy is only about −5.6 kJ/mol [24], so that adsorption of hydrogen on carbon-related materials is too weak for storing large amount of hydrogen at ambient temperature. Similar values and conclusions have also been applied to other porous materials such as zeolites [25] and metal organic frameworks (MOFs) [26].
On the other hand, if Δ H = −5.6 kJ/mol for hydrocarbons, including polymers, is used in Equation (6), the calculated optimal temperature for highest delivery Topt = 114.4 K, which is much lower than the ambient temperature needed to meet the DOE specifications. This analysis explains why most highly porous materials, including polymers, can adsorb a relatively large amount of H2 at liquid nitrogen temperature (77K), some of them up to 8.0 wt %, but the capacity drops to <1.0 wt % under ambient conditions.
Therefore, to increase H2 adsorption ability at room temperature via physisorption, we must increase the adsorption enthalpy Δ H . To achieve this, several methods from the theoretical perspective have been proposed. A first approach is to create as many ultra micropores (<1 nm) at the atomic scale as possible. Theoretical simulations show that at low pressure and high temperature (desired conditions of the DOE), graphite with an inter-layer distance of 6 Å (0.6 nm) has the highest adsorption capability, of above 10 kJ/mol, achieved through the overlap of adsorption potentials from the opposite walls [27]. In addition to such 1D slit pores, 2D cylindrical, and 3D spherical pores exhibit the same principle: cylindrical pores with r ≈ 2.0 Å and spherical pores with r ≈ 3.8 Å have maximum adsorption abilities [28,29]. Furthermore, although the free H2 molecules at room temperature behavior classically, when localized in ultra micropores, H2 molecules manifest quantum confinement behavior via the quantum sieving effect [30,31], which persists up to 300 K [26]. The quantum sieving effect versus the pore size is illustrated in Figure 1, in which the optimal pore diameter, on the order of the de Broglie wavelength for H2, corresponds to the point in which both sides of the pore interact with the absorbed molecule (corresponding to Figure 1a, blue band λH2; Figure 1b, region A; and point 2 on Figure 1c) [32].
The theoretical predictions for enhancement of H2 adsorption due to ultra micropores have been demonstrated by the experiment. For example, Gallego et al., employed in-situ small-angle neutron scattering to study H2 adsorption in activated carbons with different pore sizes, and concluded that the smaller the pore size, the larger the absorbed H2 density, as shown in Figure 2 [33]. More recently, Lee and coworkers reported that hard carbon materials with the largest micropore (<1.05 nm) volume showed the greatest H2 uptake at ambient temperature and high pressure [34]. It is of particular interest to find from the figure that the density of adsorbed H2 inside pores with 9 Å is almost same as the liquid H2 density at 298 K and 200 bar.
A second approach to improve adsorption enthalpy is to introduce charge sites in porous materials, creating charge-induced interaction between H2 and adsorbent [35]. At the molecular level, H2 molecules interact with each other or with other non-polar molecules such as hydrocarbon polymers via quadropole–quadropole interaction due to the London dispersion effect and generate a very weak van der Waals attractive force. However, when charged sites are present, additional charge-induced forces, such as dipole-induced dipole interactions, are generated, as shown in the schematic diagram Figure 3. The dependence of interactions with the distance of H2 from a host site, as well as the magnitude of interactions is shown in Table 2. The sum of the interactions determines the overall adsorption enthalpy Δ H of H2 molecules.
A third approach to increase absorption enthalpy is to create orbital interactions between an H2 molecule and d-orbitals of a transition metal intercalated in the highly porous materials [36]. The typical adsorption enthalpy from orbital interactions is 20–160 kJ/mol and therefore exceeds the optimal value for the maximum delivery of H2. This adsorption of H2 is on the order of a chemical adsorption mechanism. Such metals are typically intercalated in covalent organic frameworks (COFs), which are technically classified as molecular crystals rather than polymeric materials, and hence are beyond the scope of this review.
Although the adsorption enthalpy is the key parameter needed to retain H2 molecules at room temperature, the internal surface area of micropores is another essential parameter in storing needed quantities of H2. Internal surface area is paramount, as the critical temperature of H2 (22 K) is well below the most practical cryogenic temperature afforded by liquid nitrogen temperature (77 K). Consequently, even at liquid nitrogen temperature, only a single layer of H2 molecules is adsorbed on the internal surface of micropores in highly porous materials, including polymers. This is why the Langmuir model, which is based on the single layer adsorption, can be used to obtain the optimal adsorption enthalpy Δ H . The presumption of the Langmuir model is that adsorption ability is proportional to the number of active adsorbing sites, which are in turn proportional to the total surface area of the internal wall of micropores. Therefore, most published reports for H2 storage polymers optimize high internal surface area through synthetic strategies and/or physical methods. The corresponding H2 adsorption ability is measured at liquid nitrogen temperature (77 K), not ambient temperatures needed for practical use. In the following section, we review different classes of highly porous polymers with large internal surface areas, achieved by employing different synthetic methods and by adding different functional groups.

3. Highly Porous Organic Polymers for H2 Storage

Several types of highly porous materials for H2 storage, other than organic polymers have been explored, with the aim of meeting the DOE criteria. They include: (a) Activated carbon and its modifications: a variety of methods were used to generate different types of activated carbon with large internal surface areas [37], as well as doping the carbons with different elements such as Pt, Pd, Rh, Ni, and Cu to enhance the adsorption enthalpy [38]. Surface area of some specially-activated carbons, e.g. open carbon frameworks (OCFs), can reach as high as 3800–6500 m2/g and excess H2 adsorption ability can reach 8.5 wt % at 77 K and 100 bar [39]; (b) Carbon nanotubes (CNTs): many studies have been performed on H2 storage in single walled (SWCNTs) or multiwalled carbon nanotubes (MWCNTs). The H2 storage capacity in CNTs, particularly SWCNTs with some metal doping, is relatively high, around 7.0–10 wt %, at 77 K. Some reports claim 14–20 wt % of storage at the temperature of 298 K or above, but these results have not been widely replicated [40]. There are some excellent reviews on the H2 storage in CNT materials [17,41]; (c) Metal organic frameworks (MOFs): MOFs are crystalline solids composed of multidentate organic ligands connecting to metal ions. The surface area of MOFs can be extremely high, e.g., 6200 m2/g [18,42]. A wide range of MOF materials have been synthesized [43,44]. Here are just a few examples: For instance, Zn4O(CO2)6 has an adsorption of 1.0 wt % at 298 K and 20 bar and 4.5–7.5% at 78 K and 0.8 bar; MOF-5 with composition Zn4O(BDC)3 (BDC = 1,4-benzenedicarboxylate) gives 4.5 wt % at 77 K and 1.0 wt % at 298 K and 20 bar [21]. MOFs have weak interactions and readily release H2, but storage capacity is low under ambient conditions; and (d) Covalent organic frameworks (COFs): COFs are similar to MOFs, but with organic covalent bonds linking the frameworks together. The internal surface area of COFs is generally smaller than that of MOFs, however, lower mass organic elements and better stability of the networks make COFs an attractive material for H2 storage [34,45]. Similar to all other porous carbon materials, while COFs are good candidates for storing H2 gas at low temperature, their performance at ambient temperatures is low. Recently, several theoretical simulations suggest that the intercalation of other elements in COFs can enhance the H2 adsorption ability [46,47,48,49]. Comparison of H2 storage in the above porous materials has been reviewed by Liu et al., and a summary plot is shown in Figure 4 [50]. As seen in Liu’s review, and illustrated in Figure 4, MOFs outperform COFs and activated carbon at 77 K due to their high surface area. However, when at room temperature, none of them can meet the DOE criteria due to their small adsorption enthalpy. In addition, these higher storage capacities are achieved at high pressure, not 1 bar as would be needed for practical applications. A compilation of data collected at 1 bar/77K demonstrates reduced H2 storage capacity for MOFs of between 1.87-2.25 wt % [51].
All the above-mentioned materials are highly porous organic or organometallic crystals. In the remainder of this review, we will focus on amorphous, porous organic polymers. The structures of these polymers are varied, and can be tuned by varying synthetic routes and monomers. The different types of porous organic polymers and their performances are summarized as follows.

3.1. Hypercrosslinked Polymers (HCPs)

Hypercrosslinked polymers are co-polymers synthesized via the Friedel–Crafts method, e.g., poly(styrene-co-vinylbenzyl chloride) (PS-VBC) [52]. By choosing the right monomers, the polymer will retain networks with a very fine pore structure, and large internal surface area [53]. The surface area of this type of polymer can reach 2000 m2/g with pore sizes of 2–4 nm, and hydrogen storage capacity of 5 wt % at 77 K/80 bar, but only 0.2 wt % at 298 K/90 bar [54]. In addition to the high internal surface area, hypercrosslinked polymers also retain a high content of micropores. Large internal surface area and high micropore content make this type of polymer an attractive candidate for H2 storage. The basic synthetic route for hyper-cross linked polymers is shown in Figure 5. The benzene groups in polystyrene (PS) polymers are crosslinked by the crosslinkers, such as RCICClR’, and the microporosity within depends on the precursors used in the synthesis. When the initial precursor in making hypercrosslinked polymers is linear polystyrene (PS), as shown in Figure 5, the surface area can reach 1000 m2/g [55,56].
In addition to PS, other precursors have been used to create higher internal surface area; for instance, when Poly(vinylbenzyl chloride)-co-DVB (VBC-DVB) was used as the precursor, the internal surface area was extended to 1900 cm2/g, and the corresponding H2 storage ability at 77 K/15bar could reach 3.0 wt % [53]. Other precursors, such as polyanilines [57], polypyrroles [58], bischloromethyl monomers [59], etc. were also explored. The surface area for hypercrosslinked polymers based on these precursors can reach 632 m2/g, and the H2 storage ability at 77 K/30 bar reached 2.2 wt %. These values are significantly lower that the best MOFs and COFs a 77 K and high pressure, both of which possess much larger internal surface area. This is understandable since at 77 K, the surface area is the key parameter needed to store large amounts of H2 molecules. However, as the temperature increases to room temperature, both internal surface area and adsorption enthalpy play vital roles in storing and delivering the large amounts of H2 needed to meet DOE specifications, and hypercrosslinked polymers with more micropores could be competitive. In addition, at 1 bar pressure better-performing HPC’s have comparable H2 storage capacities to MOFs (see Table 3). Some typical hypercrosslinked polymers (HCPs) and their specific surface areas (SSAs) as well as H2 adsorption capacities are listed in Table 3.
To enhance the gas adsorption, some metals, or metal complexes such as ferrocene, have been incorporated into hypercrosslinked polymers [60,61]. However, H2 storage ability did not improve significantly for these systems at 77K, remaining at about 1.0 wt %, probably due to the continued small internal surface area of about 1000 m2/g. In addition, other elements, such as Si, have also been included in the polymer structures to improve the thermal stability. However, there are only small changes in the surface area (~1200 m2/g) and H2 storage capacity (~1.25 wt %) at 77 K/1.12 bar [62,63]. The intercalated sulfur atoms in hypercrosslinked pitch samples provide incrementally better performance, with internal surfaces areas of 1377 m2/g, yielding 1.83 wt % H2 storage at 77 K/1.13 bar [64].
Most hypercrosslinked polymers are synthesized using metal catalysts. In contrast, hydroxy-group-containing porous organic polymers have been synthesized using organic catalysts. However, both surface area (up to 920 m2/g) and H2 adsorption capacity (up to 1.28 wt % at 77 K/1 bar) remain similar to other hydrocarbon hypercrosslinked polymers [73].
While most hypercrosslinked polymers were prepared using aromatic precursors, recently other hypercrosslinked polymers were derived from chlorinated polypropylene (CPP) grafted with polyethylenimine (PEI) via a hydrothermal amination reaction [74]. Different types of CPP-g-PEI co-polymers were synthesized. The H2 adsorption capacity was reported as high as 11.26 wt % at 77 K/50 bar and 2.47 wt % at 300 K/50 bar measured with commercial H2 storage analyzer (FineSorb-3110,) at 77K and 300 K. The H2 adsorption enthalpy of these CPP-g-PEI copolymers was calculated to be 38.79 kJ/mol, indicating that chemical, rather than physical adsorption may be in play. Thus, there might be a significant barrier to releasing H2 gas under ambient conditions. In contrast, the hypercrosslinked polymers that undergo physioadsorption show much lower H2 uptake at high pressure (3 wt % or less) on HCPs than on the best MOFs and COFs (up to 10 wt % at high pressure, as indicated in Section 2).

3.2. Polymers of Intrinsic Microporosity (PIMs)

McKeown et al. initially developed PIMs by crosslinking close planar phthalocyanine macrocycles, which retain high surface area and contain rigid and nonlinear linkers preventing stacking of the monomers. The original crosslinker was derived from an agent, 5,59,6,69-tetrahydroxy-3,3,39,39-tetramethyl-1,19-spirobisindane, to make PIM-1 as shown in Figure 6 [75]. This coupling takes place under exceptionally mild reaction conditions, and in the absence of a transition metal. Other types of monomers and crosslinkers were also tried. However, it was found the PIM-1 retained the highest internal surface area, up to 1000 m2/g, and H2 adsorption ability was close to 1.7 wt % at 77 K and 10 bar. The high performance of PIM-1 has led to additional studies including aging and high pressures. For example, in a recent study Rochat et al. demonstrated that PIM-1 is stable over 400 days, showing only modest loss of hydrogen storage capacity, from 2.60 wt % (77 K/100 bar) on day 1 to 1.90 wt % on day 400 [76].
Using the same synthetic strategy, other types of PIMs were also synthesized, such as those composed of rigid and distorted macromolecules with fused-ring components, as shown in Figure 7. The Brunauer–Emmett–Teller (BET) surface (SBET) area, based on the BET model for N2 absorption, is 830 m2/g and the H2 adsorption ability at 77 K and 1 bar is ~1.43 wt % for cyclotricatechylene CTC-PIM polymers of Figure 7a [77]. Other PIMs based on same reaction mechanism with the similar monomers, and similar mild conditions for synthesis such as PIM-7, HATN-PIM, and Porph-PIM, were also studied by McKeown et al. [78,79]. These investigators found that, for similar structures, both SBET, and the H2 uptake ability are also similar. One PIM had a distinctly different structure made of non-planar units is Trip(R)-PIM (Figure 7b). This polymer results from a synthesis of triptycene subunits prepared through a dibenzodioxane formation reaction of hexahydroxyltriptycene and tetrafluoroterephthalonitrile [80]. In this system, the surface area is much higher, ~1760 m2/g, and the corresponding H2 uptake capacity is 1.79 wt % at 77 K and 1 bar. Recently, a set of three triptycene based microporous polymers (TMPs), created through Friedel–Crafts alkylations between tryptocene and multi-bromomethyl substituted benzenes were prepared. These systems share the rigid backbones of PIMS with flexible benzylic bonds. The best performing polymer, TMP-3, exceeded the performance of other PIMS, with a total surface area (SBET = 1372 m2/g), significant micropore volume (0.163 mL/g), and 4.42 wt % H2 uptake at 77 K/1 bar [81]. The advantage of these three systems two PIMs is that all three show ultra microporosity with the range of 6–8 Å, which is needed to retain H2 molecules at room temperature.
In addition to PIMs synthesized via the dibenzodioxane reaction, other PIMs were made via imidization and amidization, using amine or amide precursors, as shown in schematic reaction diagrams in Figure 8 [82,83]. The synthesis of PIMs and its applications was systematically reviewed by Ramimoghadam et al. [84]. Generally speaking, polymer chains in PIMs must contain aromatic heterocyclic ladder components in the polymer backbone that restricts the free rotation of the backbone, preventing dense packing, and therefore retaining high porosity.
Different monomers are used to generate different PIM polymers. The complete list of monomers for all three types of reactions can be found in [84]. Here we give an overview of the PIM polymers produced by these three synthetic pathways, in terms of the surface area and hydrogen storage capacity. From the general structures of the products in Figure 8, it is seen the first PIM is most rigid due to the restriction of rotation by two –O– bonds in the backbone, while the third is least rigid due to the free rotation of –CO– and –NH– bonds. The corresponding internal surface area of the first type of PIM is 120–1760 m2/g, and that of PIM–PIs is 551–1407 m2/g and that of PIM–PAs is 50–156 m2/g, with surface area decreasing with increasing bond mobility. Therefore, the first type of synthesis is more interesting in terms of hydrogen storage and a variety of different monomers have been used to make PIM polymers of this type for H2 storage. Since hydrogen storage is more about meeting technical specifications, rather than methodology development, we hereby highlight the PIMs with highest surface area. The PIMs with the largest surface area are called star triptycene-based microporous polymer (STPs), which have structures similar to Trip(R)-PIM as in Figure 7b. STPs can reach as high surface areas as high as 2000 m2/g and H2 storage capacity at 77 K and 1 bar can reach 1.92 wt %. Therefore, PIMs with star-shaped structures possess higher specific surface area (SSA) and then higher H2 uptake, aided by their three-dimensional networks. As seem with HCPs, the best PIMs still fall significantly short of H2 adsorption achieved with the best MOFs at elevated pressures; however, PIMs outperform MOFs at 1 bar/77K, with H2 storage capacities up to 4.5 wt % vs. MOFs of > 2.5%. Both fall short of DOE targets.

3.3. Conjugated Microporous Polymers (CMPs)

Since the first discovery of conjugated microporous polymers (CMPs) by Cooper et al. [85,86], these polymers have been widely applied, including for H2 gas storage. The CMPs are amorphous polymers that are made of rigid blocks linked via π-conjugated bonds, and possess three-dimensional (3D) network structures, as shown in Figure 9. The diversity of building blocks and a wide availability of different reaction types, enable chemists to fine tune the microporosity in CMPs. The basic synthetic route for building 3D CMPs is the cross coupling of two different monomers, at least one of which has more than two reactive sites, as illustrated in Figure 9.
The pore size can be varied by modifying the length of a linker between the molecules with three or more branching sites, as seen for the phenylethynylene-based CMPs in Figure 10. It is interesting to note that the longer the linker, the smaller the total surface area; e.g., CMP-0 has a surface area of ~1000 m2/g and CMP-1 and -2 have surface areas of ~ 34 and 634 m2/g, respectively.
Different phenyl monomers were used to synthesize CMPs with different specific surface areas (SSA) and pore size. For instance, Jiang et al. used 1,3,5-triethynylbenzene to make HCMP-1 with SBET = 842 m2/g and pore size = 1.1–1.6 nm, and 1,4-diethynylbenzene to make HCMP-2 with SBET = 827 m2/g and pore size = 0.9–1.6 nm [86]. Yuan et al. used the similar monomers, but a different metal catalyst, dicobalt carbonyl, to generate a serious of CMPs, POP-1, POP-2, POP-3, and POP-4, with smaller pore sizes (0.7–0.9 nm) but large surface areas (SBET > 1000 m2/g) [87]. It is interesting to compare these series of CMPs in terms of H2 at 77 K as shown in Table 4. POP-3 has the largest SBET, so its H2 uptake is highest at this temperature even though the pore size is also largest, indicating that at low temperature it is surface area dominates the H2 adsorption ability.
On the other hand, POP-1 and POP-4 have the similar surface areas, but POP-1 has a higher H2 uptake due to the smaller pore size, even though the micropore volume is also smaller indicating the number of pores is about the same for the two systems. Therefore, when at 77 K, the specific surface area, rather than the pore size, contributes more to the H2 uptake. This is why most research groups have tried to make porous materials with as large surface area as possible. Unfortunately, there is no H2 adsorption data for this group of CMPs at room temperature and therefore, no comparison of H2 uptake for them under ambient conditions, when pore size would be expected to play a larger role.
Other polyaromatic-based CMPs, such as hexaphenylbenzene (HPB)-based porous organic polymers (HPOPs) [88], tetraphenylethylene (TPE)-based porous organic polymers (TPOPs) [89,90], phenolic-resin porous organic polymers (PPOPs) [73], and carbazole-spacer-carbazole type conjugated microporous polymers (P-1 and P-2) [91] have also been prepared and tested for H2 storage characteristics. The highest H2 uptake for HPOPs is HPOP-1 (SBET = 1148 m2/g) with 1.50 wt % at 77 K/1.13 bar, for TPOPs (SBET = 810 m2/g) is TPOP-5 with 1.07 wt % at 77 K/1.13 bar, for PPOPs is PPOP-3 (SBET = 880 m2/g) with 1.28 wt % at 77 K/1 bar and for P-1/2 is P-2 (SBET = 1222 m2/g) with 1.66 wt % at 77 k/1 bar. There is no significant improvement in terms of H2 adsorption for these different modified CMPs, and all fall far short of DOE specifications.
Recently, three novel CMPs were synthesized, based on a 4,4-difluoro-4-bora-3a, 4a-diaza-s-indacene (BODIPY) core coupled with 1,3,5-triethynylbenzene, named BDT1-3. The charged non-metal sites, in addition to high microporosity lead to improved H2 adsorption, particularly for BDT3, which had a total surface area of 101 m2/g, an average pore size of 0.7 nm, and 2.2 wt % uptake at 77K. While these values need still to be improved, presence of charged non-metals on the polymer backbone and small pore size of this system are promising [92].
In addition to varying the structure of the monomers, the microporosity can also be controlled by varying the percentages of building block precursors, as well as experimental conditions. For instance, conjugated carbazole backbone was spaced and linked by different length of alkenes to modify the pore structures. However, the BET surfaces were very similar (~800 m2/g) and the highest H2 uptake is 1.33 wt % at 1.0 bar/77 K [93]. Most pure organic CMPs possess internal surface of ~ 1000 m2/g or less, and the H2 uptake is around 1.0 wt % at ~77 K/1 bar [94,95,96]. In contrast, the Li doped CMP, with a relatively low surface area of 795 m2/g, has a much higher H2 uptake (~6.1 wt % at 77K/1 bar) than pure organic CMPs [97], due to the charge-induced dipole interaction between H2 and Li atoms as described in the Background Considerations section, and hence higher adsorption enthalpy. Unfortunately, Li ions tend to aggregate in the CMPs. As an improvement, another format of Li+, methyllthium (MeLi) doped in a naphthyl-containing conjugated microporous polymer (N-CMP) can bring excess H2 uptake to 6.5 wt % at 77 K/80 bar [98]. Halides, specifically Cl and Br have been incorporated into the amine-based CMPs, borazine-linked polymers (BLPs): BLP(Cl) and BLP(Br). BLP-2(Cl) to yield specific surface areas of up to 1174 m2/g and H2 uptake of 1.30 wt % at 77 K/1 bar, while BLP-2(Br) has specific surface area of 849 m2/g an H2 uptake of 0.98 wt % at the same conditions [99]. Other organic heteroatomic systems are tried as well. For instance, nitrogen-rich conjugated microporous polymers (N-CMPs) and nitrogen-rich azo-bridged porphyrin conjugated microporous networks (N-Azos) have been synthesized and yielded SBET up to 485 and 675 m2/g with H2 uptake of 1.02 and 1.15 wt % respectively at 77K/1bar [100,101]. Thiophene-based conjugated microporous polymers (ThPOPs) have been synthesized and tested for H2 storage [102]. Among them, ThPOP-5 that retains both high BET surface (SBET = 1300 m2/g) and micropores (Vmicro = 0.28 cm3/g) can uptake H2 gas as high as 2.17 wt % at 77 K/1 bar. Incorporating organic groups such as azo and thiophene, increases the polarity of adsorption site, in addition to modifying the porosity. However as seen in Table 2, the adsorption enthalpy of the dipole-induced dipole (~0.6 kJ/mol) is much smaller than those of the charge-induced dipole (~6.8 kJ/mol) and the charge-induced quadropole (~3.5 kJ/mol). Therefore, it seems like that Li+ ions enhance H2 storage capacity more significantly than other main group elements investigated, with H2 storage capacity for Li+ ion systems exceeding the performance of the MOFs at 77 K/1 bar.

3.4. Porous Aromatic Frameworks (PAFs)

A major breakthrough in the use of highly porous organic polymers for H2 storage, was the discovery of a porous aromatic framework (PAF-1), which was synthesized via a cross coupling reaction of the tetrahedral building block tetrakis(4-bromophenyl)methane as shown in Figure 11.
The PAF-1 retains the highly porous networks that characterize MOFs and COFs, but with an amorphous structure, so it is classified as a polymer. PAF-1 not only has a very high internal surface area (~5600 m2/g), comparable to MOFs, but also is very thermally and hydrothermally stable, similar to COFs. Therefore, it combines the advantages of both MOFs and COFs. Due to its high specific surface area, the absolute and excess H2 uptake ability for PAF-1 can reach as high as 10.7 wt % and 7.0 wt %, respectively, at 77 K and 48 bar [103]. More recently, other larger tetrahedral molecules with different building blocks have been used to synthesize new PAFs, also prepared via Yamamoto homocoupling reactions [104,105]. For instance, Yuan et al. developed a series of PAFs, porous polymer networks (PPNs), using tetrahedral precursors, X(C6H4Br)4, where X = C, Ge, Si, etc. [105]. Among them, the PPN-4, which was developed from X = Si, has internal surface as high as 6461 m2/g, comparable to that of MOFs and COFs. The excess H2 adsorption can go as high as 9.1 wt % at 77 K and 55 bar. There is no H2 adsorption data reported at 298 K. However, H2 absorption is expected to drop to a low value at ambient temperature, due to the very weak van der Waals interaction between H2 molecules and hydrocarbon network. Another approach along this line is that organic linkers with different lengths were inserted between the tetrahedral building blockers to obtain different pore sizes. For example, by inserting diphenylacetylene (DPA), 1,4-diphenyl-buta- diyne (DPB), 1,4-bis (phenylethynyl) benzene (BPEB), or 1,4-bis (phe-nylbutadiynyl) benzene (BPBB) linkers, Wu et al. [106] synthesized a series of porous aromatic frameworks (PAF-322, PAF-324, PAF-332, and PAF-334). The size of the building unit in PAFs was significantly expanded and so were the pore size and internal surface area. The total H2 uptake can go as high as 63.96 wt % at 77 K/100 bar, and excess H2 uptake can reach 10.69 wt % at 77 K/20 bar. In particular, the total H2 uptake of PAF-334 at 298 K/100 bar is simulated to be 16.03 wt %. However, it is worthwhile to mention that the above results are the simulation results, not experimental results. Furthermore, due to the large open pores inside the polymer, the volumetric capacity of H2 storage for these PAFs is only approximate 9.0 g/L at 298 K/120bar, far below the DOE criteria (40 g/L). Similarly, theoretical simulations on designing up to 115 organic PAF-XXXs show that the weight storage capability of H2 can reach 5.9 wt % at 298 K/100 bar, while the volumetric capacity of H2 storage can only reach 7.9 g/L [107]. Therefore, simply increasing the pore size via different synthetic strategies might not be the right direction to make highly porous polymers that meet the DOE standards for H2 storage.
In additional studies, several metals have been added to PAFs to enhance their H2 adsorption capacity. Similar to that in conjugated microporous polymers (CMPs), doping PAF-1 with 5% of Li (Li@PAF-1) significantly increases ultra micropore content (<10Å size) and possible adsorption enthalpy, and hence increases H2 adsorption ability by 22% [108]. Lithium ion doping in the PAF derivative, PAF-18-OH, also increases the polymer’s H2 adsorption [109]. In a related simulation, lithium-decorated fullerenes (Li6C60) were impregnated in PAFs and H2 uptake was predicted to be 5.5 wt % at 77 K/1 bar [110]. Addition of magnesium alkoxide was also studied by simulation, and is predicted to enhance the interaction between H2 and polymer pores, yielding a predicted H2 uptake of 7.12 wt % at 298 K and 100 bar [111]. Note that this is approaching the DOE 2020 targets for room temperature performance. Similar to what was observed for CMPs, a nitrogen-rich porous aromatic framework (N-PAF) was also developed with the SBET = 1790 m2/g and H2 uptake up to 1.87 wt % at 77 K/1 bar [112]. Based on this study, the nitrogen substitution does not improve material performance.
Recently, Rochat et al. copolymerized the porous aromatic framework, PAF-1, with the polymer with intrinsic microporosity, PIM-1 [113]. By varying the relative contents of PAF-1 and PIM-1, they demonstrated that the PAF polymer retains a much larger internal surface area and hence a higher H2 store capacity. In fact, PAFs are the only polymer to date to show significantly higher H2 storage capacity than the most MOFs at high pressures. Data for performance at 1 bar is limited to a single study, showing H2 storage capacity similar to MOFs under these conditions.

3.5. Other Porous Polymers

Several emerging polymeric systems for H2 storage do not fit into the four classes above, including: (1) coordination polymers, (2) polypropylene gels, and (3) phosphorous-organic polymers. Coordination polymers can provide for both pores need for H2 storage and transition metals to enhance binding through d-orbitals. In a recent paper by Lyu et al. PCP-31 and -32 were created with chelated Cu2+ open metal sites incorporated into mesopores of 2.3–2.8 nm; despite the larger pore size, up to 10 wt % of H2 at 77 k/100 bar was adsorbed for the best performing system [114]. Polyphenylene gels are similar to other aromatic polymers, but lack non-aromatic linkers. A recent study demonstrated total surface areas of 219–674 m2/g for the polyphenylenes prepared, with assumed high microporosity. However, H2 adsorption data was not reported [115]. Ahmed et al. prepared three branched organic phosphate esters with azo linkages; despite low total surface values (up to 30.0 m2/g) and low total pore volume (up to 0.052 cm3/g), H2 adsorption at near ambient temperature (323 K/50 bar) is moderate (up to 0.66 wt %), signaling room for improvement with further structural modification [116].

3.6. Brief comparison of highly porous organic polymers with MOFs

To compare the performance of different materials in H2 storage, the experimental conditions for measuring the H2 uptake should be same, or at least similar, that is, under the same temperature and pressure. Unfortunately, the reported results available were obtained at different conditions. Although the temperatures for the experiments were at either 77 K or room temperature, the pressures for H2 measurement varied significantly, ranging from 1 bar to hundreds of bar. Such discrepancy in the experimental conditional makes the direct comparison impossible. Nevertheless, we selected the results under the similar conditions, close to 77 K/1bar, to make a rough comparison, as shown in Figure 12. Since the pressure is not exactly same, e.g. some 1 bar, and other 1.13 bar as seen in Table 3, we present the general ranges of pore size vs. H2 uptake, rather than individual definite data points in the figure.
It is interesting to see that polymers with intrinsic microporosity (PIMs) outperform the other highly porous polymers in H2 adsorption at ambient pressures. PIMS even outperform MOFs when under these conditions, despite the latter ones possess much larger internal surface and pore volumes. This analysis clearly indicates that the number of ultra micropores determines the total H2 adsorption, which is very consistent with the theoretical predictions.

4. Characterization of Porosity in Highly Porous Organic Polymers

As described above, ultra microporosity and adsorption enthalpy are key parameters for a highly porous organic polymer to adsorb large amounts of H2. The adsorption enthalpy can be measured by the adsorption dynamics. The porosity is traditionally characterized by the N2 gas adsorption method owing to the high relatively critical temperature of N2 gas, and well-defined theory. Normally, N2 gas is applied at different relative pressures from 10–8 to 1 to provide an adsorption isotherm, which describes the adsorption of N2 molecules over a wide range of porosity. To extract the information on porosity from H2 adsorption isotherms, different models have been proposed based on the mechanism of the micro-filling process in pores. The most widely used model is the Brunauer–Emmett–Teller (BET) model, which takes into consideration the multilayer adsorption of N2 molecules at liquid N2 temperature [117,118], yielding the total surface area. Alternatively, the Horvathe–Kawazoe (HK) method [119,120] and Dubinin–Radushkevich (DR) analyses [121,122] are often used to analyze micropore (r < 20 Å) volume and pore distribution at low pressure with the aid of the density functional theory (DFT) [123,124,125]. All three methods are based on the same experimental data: gas adsorption—normally, N2 gas adsorption at 77 K.
Unfortunately, N2 adsorption experiments have some intrinsic limitations when measuring ultra micropores (~Å) in highly porous polymers for H2 storage at ambient conditions. First, N2 adsorption methods do not effectively measure ultra micropores that are crucial to H2 storage due to large kinetic diameter of N2 molecules, 3.64 Å [126], which is comparable to the optimum pore sizes (~3.8 Å) for maximum H2 adsorption at room temperature. The adsorption of other gas media, such as CO2 and H2 has also been used to study the porosity [127]. However, no significant improvement in measuring micropores has been achieved, due to either the large size of CO2 molecules (3.3 Å), or the low critical temperature of H2 (32 K). Second, the N2 gas adsorption method provides isotherms at liquid N2 temperature (77 K), and is unable to generate the data needed to examine the H2 storage ability of a material at room temperature. Since pore volume in polymers significantly changes with temperature, the results obtained at 77 K might be very different from those at room temperature. In particular, ultra micropores (~Å) are much more sensitive to temperature, in that the expansion coefficient of those pores is almost ten times larger than the bulk value [128]. Therefore, some candidate materials ruled out by the N2 adsorption method at 77 K may be appropriate for H2 storage at room temperature, and vice versa. Finally, the adsorption of N2 at 77 K can cause a swelling effect or warp formation in the pore structure, giving hysteresis in the adsorption-desorption isotherm curve and therefore inaccurate information on the pristine pore sizes.
Recently, another technique, positron annihilation lifetime spectroscopy (PALS) has been applied to study ultra micropores in H2 storage materials, providing a useful alternative to the N2 gas absorption method [84,108,129]. The positron is a particle with a positive charge and the same mass as an electron, and is generated from a positron source, normally a radioactive isotope 22Na. The schematic diagram for PALS technique and its mechanism is illustrated in Figure 13.
Figure 13a shows the general outline of the PALS setup. When a positron e+ emits from the radioactive source, an accompany ν ray with the energy of 1274 keV is given off simultaneously and detected by one detector, marking the birth of a positron. Then the positron will quickly slow down (at picosecond scale) and form a positronium (Ps). After diffusing and residing in a pore for a few nanoseconds, Ps will annihilate with the electron layer at the internal wall due to the overlap of Ps wave function with the above electron layer as shown in Figure 13b. The annihilation of Ps will give off another ν ray with the energy of 511 keV and detected by a second detector, marking the end of a positron. The time difference between these events is basically the lifetime of Ps, which is converted to the electronic signal via a time-amplitude-convertor (TAC) and recorded in a computer. The relationship between the lifetime of Ps and pore radius (r) is well described by the Tao–Eldrup equation as follows:
τ = 0.5 n s ( 1 R R + Δ R + s i n 2 π R R + Δ R ) 1 ,
where τ is lifetime to Ps, R is the radius of a pore, and ΔR is an empirical value, 1.66Å, representing the thickness of electron layer on the internal wall of the pore [130].
The PALS technique is able to overcome the intrinsic limitations associated with the N2 adsorption method. First, Ps has same size as a hydrogen atom, with diameter of 1.06 Å, and is particularly sensitive to the pores with radii from 2–10 Å. This is the range of ultra micropores that theory predicts to be optimal in adsorbing H2 molecules at room temperature. Second, PALS measurements can be carried out at any temperature and hence can measure the porosity of highly porous polymers at room temperature. In addition, PALS can be performed in situ with H2 adsorption/desorption processes, thus providing dynamic information on H2 adsorption. Finally, Ps themselves will not create any swelling effect since both the positron and electron constituting Ps are leptons, and it is the quantum effect, rather than the classical space filling effect, that is used to obtain the pore size in PALS.
Other techniques have also been used to study the porosity in porous polymers. For instance, 129Xe NMR spectroscopy, however, it was found Xe atoms exclusively occupy in large pores (>20 Å), not the micropores of interest [131]. Small angle X-ray scattering (SAXS) in combination with the N2 gas adsorption gives the internal surface area, however, this method still relies on the N2 adsorption isotherms with the limitations given above [83].
In summary, the technique best suited for characterizing ultra microporosity in highly porous organic polymers particularly for H2 gas storage is positron annihilation lifetime spectroscopy (PALS). However, PALS uses radioactive material, 22Na, which might limit its application even though its reactivity is relatively low ~30 μCi, and expertise with this method is not widespread.

5. Conclusions and Prospective Future

Porous organic polymers have their own advantages in H2 storage, for instance, the structures and porosity can be controlled by the use of different monomers and synthetic routes; the cost is relatively low and it is suitable for mass production, and normally, there is no heavy-metal incorporated, making the materials and processes friendly to the environment. These types of polymers store H2 molecules via physical adsorption through van der Waals interaction, and hence can release H2 as fuel under relatively mild conditions. Four types of highly porous organic polymers, hypercrosslinked polymers (HCPs), polymers with intrinsic microporosity (PIMs), conjugated microporous polymers (CMPs), and porous aromatic frameworks (PAFs) demonstrated high internal surface areas and good H2 uptake at cryogenic temperature. It is seen that the polymers originated from the star shaped precursors, such as Trip(R)-PIM and PAFs, can retain the highest specific surface area and outperform the other polymers generated from planar shaped precursors, and therefore, induce higher H2 uptake capacity under higher pressures. With the limited data published, only PIMs have been shown to have H2 storage capacities approaching 5 wt % at 77 K. However, due to the weak van der Waals interactions, the H2 adsorption enthalpy is only approximate ~6.0 kJ/mol for most carbon based polymers, and hence the H2 storage ability is very low under ambient temperature. Most of these materials would have an H2 absorption of ~1.0 wt % at room temperature, which is far below the targets specified by the DOE, despite of the fact that many highly porous polymers can store up to 10 wt % H2 at liquid N2 temperature (77 K) and high pressures (10 bar or more).
To enhance the H2 adsorption ability at room temperature and ambient pressure, it is imperative to increase H2 adsorption enthalpy, in addition to high porosity. Theoretical considerations revealed two effective methods to achieve such goals: first, producing as many ultra micropores at several Å as possible; second, introducing some light metal ions, such as Li and Na ions, to create some charge-induced dipole interactions between H2 molecules and charge site. Ultra micropores, in contrast with macropores at nm scale or mesopores at μm scale, promote interaction between H2 molecules and pore walls, via the overlap of potentials from opposite walls and the quantum sieving effect that can be maintained up to 300 K, while larger pores can only have van der Waals forces between pore walls and H2. Unfortunately, most reports give only the content of overall porosity, which is probably composed primarily of macropores rather than ultra micropores, and therefore, the systems studied yield good to excellent adsorption result at 77 K, but do not perform well at room temperature. Nevertheless, there are some reports of highly porous polymers with high ultra micropore content. For instance, Zhang et al. [132] synthesized microporous polymer HTP-B using a hexaphenylbenzene-based triptycene monomer, in an attempt to introduce some ultra micropores under 10 Å. Although the exact ultra micropore content is not known, the H2 uptake is significantly higher than that of its counterpart, HTP-A that has no ultra-micropore content (1.09 wt % vs. 0.55 wt %). Therefore, future efforts should be directed to synthesizing highly porous polymers that have both high BET surface area and high content of ultra micropores. To further enhance the room temperature H2 absorption activity of these polymers, light metal ions can be doped into the materials. A third approach might be to add some transition metals to create so called spillover of H2 molecules. Spillover is a process by which H2 molecules dissociate and bond with C atom in sp2 hybrid orbital when catalyzed by some transitions metals, such as Pt2+ [133]. The spillover mechanism for H2 storage was initially hailed by many scientists since it was reported that the adsorption enthalpy of spillover could reach 10–30 kJ/mol, and the H2 adsorption could potentially exceed both 5.5 wt % and 50 g/L that were very close to the DOE criteria [134,135,136]. However, both theoretical and experimental study later confirmed that the spillover mechanism is not sufficient enough to generate onboard H2 storage [137]. Other concerns of spillover include the quick plague of the metal catalyst due to the oxidation reaction and irreversible hydrogenation [138] and the right position of the metal ions due to the amorphous structure of porous organic polymers. Therefore, it seems like spillover is not the right approach in this type of materials for H2 storage.
As a final note, the majority of the studies to date have focused on attaining the highest possible H2 capacities by using low temperature and/or high pressure conditions. These reviewers suggest that future studies would be better served in measuring H2 storage capacities at ambient temperature and pressure. This would not only make comparing data from different studies more uniform, it would also clearly illustrate progress toward achieving DOE H2 storage goals.

Author Contributions

R.Z. initiated the work upon the invitation and formulated the overall structure with preliminary draft. He has characterized micropores in polymeric materials for over 20 years. K.C. contributed significantly to summarizing the work, adding updated references, and addressing reviewers’ comments. Her research areas span physical organic chemistry and computational materials science.

Funding

This research was funded by NATIONAL SCIENCE FOUNDATION (NSF) with grant number “1345163”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van den Berg, A.W.C.; Areán, C.O. Materials for Hydrogen Storage: Current Research Trends and Perspectives. Chem. Commun. 2008, 668–681. [Google Scholar] [CrossRef]
  2. Schlapbach, L.; Zutell, A. Hydrogen Storage Materials for Mobile Applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef]
  3. Jena, P. Materials for Hydrogen Storage: Past, Present, and Future. J. Phys. Chem. Lett. 2011, 2, 206–211. [Google Scholar] [CrossRef]
  4. Stetson, N.T. Hydrogen Storage—2010 Annual Merit Review and Peer Evaluation Meeting. 2010. Available online: http://www.hydrogen.energy.gov/pdfs/review10/st00a_stetson_2010_o_web.pdf (accessed on 31 January 2019).
  5. Satyapal, S.; Petrovic, J.; Read, C.; Thomas, G.; Ordaz, G. The U.S. Department of Energy’s National Hydrogen Storage Project: Progress towards Meeting Hydrogen-Powered Vehicle Requirements. Catal. Today 2007, 120, 246–256. [Google Scholar] [CrossRef]
  6. 2015 Hydrogen Storage. Available online: https://www.energy.gov/sites/prod/files/2015/05/f22/fcto_myrdd_storage.pdf (accessed on 31 January 2019).
  7. Target Explanation Document: Hydrogen Storage for Light-Duty Fuel Cell Vehicles 2017. Available online: https://www.energy.gov/sites/prod/files/2017/05/f34/fcto_targets_onboard_hydro_storage_explanation.pdf (accessed on 31 January 2019).
  8. Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M.; Dogan, B. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy 2007, 32, 1121–1240. [Google Scholar] [CrossRef]
  9. Amica, G.; Larochette, P.A.; Gennari, F.C. Hydrogen storage properties of LiNH2–LiH system with MgH2, CaH2 and TiH2 added. Int. J. Hydrogen Energy 2015, 40, 9335–9346. [Google Scholar] [CrossRef]
  10. Xiong, R.; Sang, G.; Zhang, G.; Yan, X.; Li, P.; Yao, Y.; Luo, D.; Chen, C.A.; Tang, T. Evolution of the active species and catalytic mechanism of Ti doped NaAlH4 for hydrogen storage. Int. J. Hydrogen Energy 2017, 42, 6088–6095. [Google Scholar] [CrossRef]
  11. Petit, J.F.; Miele, P.; Demirci, U.B. Ammonia borane H3NBH3 for solid-state chemical hydrogen storage: Different samples with different thermal behaviors. Int. J. Hydrogen Energy 2016, 41, 15462–15470. [Google Scholar] [CrossRef]
  12. Zubizarreta, L.; Arenillas, A.; Pis, J.J. Carbon materials for H2 storage. Int. J. Hydrogen Energy 2009, 34, 4575–4581. [Google Scholar] [CrossRef]
  13. Gogotsi, Y.; Dash, R.K.; Yushin, G.; Yildirim, T.; Laudisio, G.; Fischer, J.E. Tailoring of Nanoscale Porosity in Carbide-Derived Carbons for Hydrogen Storage. J. Am. Chem. Soc. 2005, 127, 16006–16007. [Google Scholar] [CrossRef] [PubMed]
  14. Benard, P.; Chahine, R.; Chandonia, P.A.; Cossement, D.; Dorval-Douville, G.; Lafi, L.; Lachance, P.; Paggiaro, R.; Poirier, E. Comparison of hydrogen adsorption on nanoporous materials. J. Alloys Compd. 2007, 446, 380–384. [Google Scholar] [CrossRef]
  15. Patchkovskii, S.; Tse, J.S.; Yurchenko, S.N.; Zhechkov, L.; Heine, T.; Seifert, G. Graphene Nanostructures as Tunable Storage Media for Molecular Hydrogen. Proc. Natl. Acad. Sci. USA 2005, 102, 10439–10444. [Google Scholar] [CrossRef] [PubMed]
  16. Jordá-Beneyto, M.; Suárez-García, F.; Lozano-Castelló, D.; Cazorla-Amorós, D.; Linares-Solano, A. Hydrogen storage on chemically activated carbons and carbon nanomaterials at high pressures. Carbon 2007, 45, 293–303. [Google Scholar] [CrossRef]
  17. Yürüm, Y.; Taralp, A.; Veziroglu, T.N. Storage of Hydrogen in Nanostructured Carbon Materials. Int. J. Hydrogen Energy 2009, 34, 3784–3798. [Google Scholar] [CrossRef]
  18. Dalebrook, A.F.; Gan, W.; Grasemann, M.; Moret, S.; Laurenczy, G. Hydrogen Storage: Beyond Conventional Methods. Chem. Commun. 2013, 49, 8735–8751. [Google Scholar] [CrossRef]
  19. Broom, D.P.; Webb, C.J.; Fanourgakis, G.S.; Froudakis, G.E.; Trikalitis, P.N.; Hirscher, M. Concepts for Improving Hydrogen Storage in Nanoporous Materials. Int. J. Hydrogen Energy 2019, 44, 7768–7779. [Google Scholar] [CrossRef]
  20. Bhatia, S.K.; Myers, A.L. Optimum conditions for adsorptive storage. Langmuir 2006, 22, 1688–1700. [Google Scholar] [CrossRef]
  21. Rosi, N.L.; Eckert, J.; Eddaoudi, M.; Vodak, D.T.; Kim, J.; O’keeffe, M.; Yaghi, O.M. Hydrogen storage in microporous metal-organic frameworks. Science 2003, 300, 1127–1129. [Google Scholar] [CrossRef]
  22. Sagara, T.; Klassen, J.; Ganz, E. Computational study of hydrogen binding by metal-organic framework-5. J. Chem. Phys. 2004, 121, 12543–12547. [Google Scholar] [CrossRef]
  23. Allendorf, M.D.; Hulvey, Z.; Gennett, T.; Ahmed, A.; Autrey, T.; Camp, J.; Cho, E.S.; Furukawa, H.; Haranczyk, M.; Head-Gordon, M.; et al. An assessment of strategies for the development of solid-state adsorbents for vehicular hydrogen storage. Energy Environ. Sci. 2018, 11, 2784–2812. [Google Scholar] [CrossRef]
  24. Heine, T.; Zhechkov, L.; Seifert, G. Hydrogen storage by physisorption on nanostructured graphite platelets. Phys. Chem. Chem. Phys. 2004, 6, 980–984. [Google Scholar] [CrossRef]
  25. Otero Areán, C.; Manoilova, O.V.; Bonelli, B.; Rodríguez Delgado, M.; Turnes Palomino, G.; Garrone, E. Thermodynamics of Hydrogen Adsorption on the Zeolite Li-ZSM-5. Chem. Phys. Lett. 2003, 370, 631–635. [Google Scholar] [CrossRef]
  26. Garberoglio, G.; Skoulidas, A.I.; Johnson, J.K. Adsorption of Gases in Metal Organic Materials:  Comparison of Simulations and Experiments. J. Phys. Chem. B 2005, 109, 13094–13103. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, Q.; Johnson, J.K. Molecular simulation of hydrogen adsorption in single-walled carbon nanotubes and idealized carbon slit pores. J. Chem. Phys. 1990, 110, 577–586. [Google Scholar] [CrossRef]
  28. Cabria, I.; Lopez, M.J.; Alonso, J.A. Simulation of the hydrogen storage in nanoporous carbons with different pore shapes. Int. J. Hydrogen Energy 2011, 36, 10748–10759. [Google Scholar] [CrossRef]
  29. Rzepka, M.; Lamp, P.; de la Casa-Lillo, M.A. Physisorption of Hydrogen on Microporous Carbon and Carbon Nanotubes. J. Phys. Chem. B 1998, 102, 10894–10898. [Google Scholar] [CrossRef] [Green Version]
  30. Wang, Q.; Challa, S.R.; Sholl, D.S.; Johnson, J.K. Quantum sieving in carbon nanotubes and zeolites. Phys. Rev. Lett. 1999, 82, 956–959. [Google Scholar] [CrossRef]
  31. Cai, J.; Xing, Y.; Zhao, X. Quantum sieving: Feasibility and challenges for the separation of hydrogen isotopes in nanoporous materials. RSC Adv. 2012, 2, 8579–8586. [Google Scholar] [CrossRef]
  32. Oh, H.; Hirscher, M. Quantum sieving for separation of hydrogen isotopes using MOFs. Eur. J. Inorg. Chem. 2016, 27, 4278–4289. [Google Scholar] [CrossRef]
  33. Gallego, N.C.; He, L.; Saha, D.; Contescu, C.I.; Melnichenko, Y.B. Hydrogen confinement in carbon nanopores: Extreme densification at ambient temperature. J. Am. Chem. Soc. 2011, 133, 13794–13797. [Google Scholar] [CrossRef] [PubMed]
  34. Lohse, M.S.; Bein, T. Covalent Organic Frameworks: Structures, Synthesis, and Applications. Adv. Funct. Mater. 2018, 28, 1705553. [Google Scholar] [CrossRef]
  35. Lochan, R.C.; Head-Gordon, M. Computational studies of molecular hydrogen binding affinities: The role of dispersion forces, electrostatics, and orbital interactions. Phys. Chem. Chem. Phys. 2006, 8, 1357–1370. [Google Scholar] [CrossRef]
  36. Gonzalez, A.A.; Zhang, K.; Nolan, S.P.; Lopez de la Vega, R.L.; Mukerjee, S.L.; Hoff, C.D.; Kubas, G.J. Thermodynamic and Kinetic Studies of the Complexes W(CO)3(PCy3)2(L) (L = H2, N2, NCCH3, Pyridine, P(OMe)3, CO). Organometallics 1988, 7, 2429–2435. [Google Scholar] [CrossRef]
  37. Sevilla, M.; Fuertes, A.B.; Mokaya, R. High Density Hydrogen Storage in Superactivated Carbons from Hydrothermally Carbonized Renewable Organic Materials. Energy Environ. Sci. 2011, 4, 1400–1410. [Google Scholar] [CrossRef]
  38. Rossetti, I.; Ramis, G.; Gallo, A.; Michele, A.D. Hydrogen Storage over Metal-Doped Activated Carbon. Int. J. Hydrogen Energy 2015, 40, 7609–7616. [Google Scholar] [CrossRef]
  39. Kuchta, B.; Firlej, L.; Mohammadhosseini, A.; Boulet, P.; Beckner, M.; Romanos, J.; Pfeifer, P. Hypothetical High-Surface-Area Carbons with Exceptional Hydrogen Storage Capacities: Open Carbon Frameworks. J. Am. Chem. Soc. 2012, 134, 15130–15137. [Google Scholar] [CrossRef] [PubMed]
  40. Lee, S.M.; Park, K.S.; Choi, Y.C.; Park, Y.S.; Bok, J.M.; Bae, D.J.; Nahm, K.S.; Choi, Y.G.; Yu, S.C.; Kim, N.; et al. Hydrogen Adsorption and Storage in Carbon Nanotubes. Synth. Met. 2000, 113, 209–216. [Google Scholar] [CrossRef]
  41. Li, F.; Zhao, J.; Chen, Z. Carbon-Based Nanomaterials for H2 Storage. In Carbon Nanomaterials for Advanced Energy Systems; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2015; pp. 407–437. [Google Scholar]
  42. Li, G.; Kobayashi, H.; Taylor, J.M.; Ikeda, R.; Kubota, Y.; Kato, K.; Takata, M.; Yamamoto, T.; Toh, S.; Matsumura, S.; et al. Hydrogen Storage in Pd Nanocrystals Covered with a Metal–Organic Framework. Nat. Mater. 2014, 13, 802. [Google Scholar] [CrossRef]
  43. Furukawa, H.; Miller, M.A.; Yaghi, O.M. Independent Verification of the Saturation Hydrogen Uptake in MOF-177 and Establishment of a Benchmark for Hydrogen Adsorption in Metal–Organic Frameworks. J. Mater. Chem. 2007, 17, 3197–3204. [Google Scholar] [CrossRef]
  44. Wong-Foy, A.G.; Matzger, A.J.; Yaghi, O.M. Exceptional H2 saturation uptake in microporous metal−organic frameworks. J. Am. Chem. Soc. 2006, 128, 3494–3495. [Google Scholar] [CrossRef]
  45. Feng, X.; Ding, X.; Jiang, D. Covalent organic frameworks. Chem. Soc. Rev. 2012, 41, 6010–6022. [Google Scholar] [CrossRef]
  46. Pramudya, Y.; Mendoza-Cortes, J.L. Design principles for high H2 storage using chelation of abundant transition metals in covalent organic frameworks for 0–700 bar at 298 K. J. Am. Chem. Soc. 2016, 138, 15204–15213. [Google Scholar] [CrossRef]
  47. Mendoza-Cortes, J.L.; Goddard, W.A.; Furukawa, H.; Yaghi, O.M. A Covalent Organic Framework That Exceeds the DOE 2015 Volumetric Target for H2 Uptake at 298 K. J. Phys. Chem. Lett. 2012, 3, 2671–2675. [Google Scholar] [CrossRef]
  48. Klontzas, E.; Tylianakis, E.; Froudakis, G.E. Hydrogen Storage in Lithium-Functionalized 3-D Covalent-Organic Framework Materials. J. Phys. Chem. C 2009, 113, 21253–21257. [Google Scholar] [CrossRef]
  49. Gao, F.; Sun, J.T.; Meng, S. “H2 Sponge”: Pressure as a Means for Reversible High-Capacity Hydrogen Storage in Nanoporous Ca-Intercalated Covalent Organic Frameworks. Nanoscale 2015, 7, 6319–6324. [Google Scholar] [CrossRef]
  50. Liu, J.; Zou, R.; Zhao, Y. Recent Developments in Porous Materials for H2 and CH4 Storage. Tetrahedron Lett. 2016, 57, 4873–4881. [Google Scholar] [CrossRef]
  51. Yuan, D.; Zhao, D.; Sun, D.; Zhou, H. An Isoreticular Series of Metal–Organic Frameworks with Dendritic Hexacarboxylate Ligands and Exceptionally High Gas-Uptake Capacity. Angew. Chem. Int. Ed. 2010, 49, 5357–5361. [Google Scholar] [CrossRef] [PubMed]
  52. Veverka, P.; Jeřábek, K. Mechanism of Hypercrosslinking of Chloromethylated Styrene–Divinylbenzene Copolymers. React. Funct. Polym. 1999, 41, 21–25. [Google Scholar] [CrossRef]
  53. Germain, J.; Hradil, J.; Fréchet, J.M.J.; Svec, F. High Surface Area Nanoporous Polymers for Reversible Hydrogen Storage. Chem. Mater. 2006, 18, 4430–4435. [Google Scholar] [CrossRef]
  54. Germain, J.; Fréchet, J.M.J.; Svec, F. Nanoporous Polymers for Hydrogen Storage. Small 2009, 5, 1098–1111. [Google Scholar] [CrossRef]
  55. Fontanals, N.; Marcé, R.M.; Cormack, P.A.G.; Sherrington, D.C.; Borrull, F. Monodisperse, Hypercrosslinked Polymer Microspheres as Tailor-Made Sorbents for Highly Efficient Solid-Phase Extractions of Polar Pollutants from Water Samples. J. Chromatogr. A 2008, 1191, 118–124. [Google Scholar] [CrossRef] [PubMed]
  56. Guiochon, G.; Gritti, F. Shell particles, trials, tribulations and triumphs. J. Chrom. A 2011, 1218, 1915–1938. [Google Scholar] [CrossRef] [PubMed]
  57. Germain, J.; Fréchet, J.M.J.; Svec, F. Hypercrosslinked Polyanilines with Nanoporous Structure and High Surface Area: Potential Adsorbents for Hydrogen Storage. J. Mater. Chem. 2007, 17, 4989–4997. [Google Scholar] [CrossRef]
  58. Germain, J.; Fréchet, J.M.J.; Svec, F. Nanoporous, Hypercrosslinked Polypyrroles: Effect of Crosslinking Moiety on Pore Size and Selective Gas Adsorption. Chem. Commun. Camb. Engl. 2009, 12, 1526–1528. [Google Scholar] [CrossRef]
  59. Wood, C.D.; Tan, B.; Trewin, A.; Niu, H.; Bradshaw, D.; Rosseinsky, M.J.; Khimyak, Y.Z.; Campbell, N.L.; Kirk, R.; Stöckel, E.; et al. Hydrogen Storage in Microporous Hypercrosslinked Organic Polymer Networks. Chem. Mater. 2007, 19, 2034–2048. [Google Scholar] [CrossRef]
  60. Gagnon-Thibault, É.; Cossement, D.; Guillet-Nicolas, R.; Masoumifard, N.; Bénard, P.; Kleitz, F.; Chahine, R.; Morin, J.F. Nanoporous Ferrocene-Based Cross-Linked Polymers and Their Hydrogen Sorption Properties. Microporous Mesoporous Mater. 2014, 188, 182–189. [Google Scholar] [CrossRef]
  61. Li, G.; Liu, Q.; Xia, B.; Huang, J.; Li, S.; Guan, Y.; Zhou, H.; Liao, B.; Zhou, Z.; Liu, B. Synthesis of Stable Metal-Containing Porous Organic Polymers for Gas Storage. Eur. Polym. J. 2017, 91, 242–247. [Google Scholar] [CrossRef]
  62. Shu, G.; Zhang, C.; Li, Y.; Jiang, J.X.; Wang, X.; Li, H.; Wang, F. Hypercrosslinked silole-containing microporous organic polymers with N-functionalized pore surfaces for gas storage and separation. J. Appl. Polym. Sci. 2018, 135, 45907. [Google Scholar] [CrossRef]
  63. Fu, S.; Yao, J.; Yang, Z.; Sun, H.; Liu, W. Silane-Based Hyper-Cross-Linked Porous Polymers and Their Applications in Gas Storage and Water Treatment. J. Mater. Sci. 2018, 53, 10469–10478. [Google Scholar] [CrossRef]
  64. Gao, H.; Ding, L.; Bai, H.; Liu, A.; Li, S.; Li, L. Pitch-based hyper-cross-linked polymers with high performance for gas adsorption. J. Mater. Chem. A 2016, 4, 16490–16498. [Google Scholar] [CrossRef]
  65. Li, B.; Huang, X.; Liang, L.; Tan, B. Synthesis of uniform microporous polymer nanoparticles and their applications for hydrogen storage. J. Mater. Chem. 2010, 20, 7444–7450. [Google Scholar] [CrossRef]
  66. Pan, L.; Chen, Q.; Zhu, J.H.; Yu, J.G.; He, Y.J.; Han, B.H. Hypercrosslinked porous polycarbazoles via one-step oxidative coupling reaction and Friedel–Crafts alkylation. Polym. Chem. 2015, 6, 2478–2487. [Google Scholar] [CrossRef]
  67. Yang, X.; Yu, M.; Zhao, Y.; Zhang, C.; Wang, X.; Jiang, J.X. Hypercrosslinked microporous polymers based on carbazole for gas storage and separation. RSC Adv. 2014, 4, 61051–61055. [Google Scholar] [CrossRef]
  68. Gao, H.; Ding, L.; Li, W.; Ma, G.; Bai, H.; Li, L. Hyper-cross-linked organic microporous polymers based on alternating copolymerization of bismaleimide. ACS Macro Lett. 2016, 5, 377–381. [Google Scholar] [CrossRef]
  69. Liu, G.; Wang, Y.; Shen, C.; Ju, Z.; Yuan, D. A facile synthesis of microporous organic polymers for efficient gas storage and separation. J. Mater. Chem. A 2015, 3, 3051–3058. [Google Scholar] [CrossRef]
  70. Li, B.; Gong, R.; Wang, W.; Huang, X.; Zhang, W.; Li, H.; Hu, C.; Tan, B. A new strategy to microporous polymers: Knitting rigid aromatic building blocks by external cross-linker. Macromolecules 2011, 44, 2410–2414. [Google Scholar] [CrossRef]
  71. Schwab, M.G.; Lennert, A.; Pahnke, J.; Jonschker, G.; Koch, M.; Senkovska, I.; Rehahn, M.; Kaskel, S. Nanoporous copolymer networks through multiple Friedel–Crafts-alkylation—Studies on hydrogen and methane storage. J. Mater. Chem. 2011, 21, 2131–2135. [Google Scholar] [CrossRef]
  72. Dawson, R.; Stevens, L.A.; Drage, T.C.; Snape, C.E.; Smith, M.W.; Adams, D.J.; Cooper, A.I. Impact of water coadsorption for carbon dioxide capture in microporous polymer sorbents. J. Am. Chem. Soc. 2012, 134, 10741–10744. [Google Scholar] [CrossRef]
  73. Wang, T.; Zhao, Y.C.; Zhang, L.M.; Cui, Y.; Zhang, C.S.; Han, B.H. Novel approach to hydroxy-group-containing porous organic polymers from bisphenol A. Beilstein J. Org. Chem. 2017, 13, 2131–2137. [Google Scholar] [CrossRef] [Green Version]
  74. Chen, S.; Liu, J.; Li, Z.; Wang, H.; Wang, X.; Xu, Y. Hydrogen storage properties of highly cross-linked polymers derived from chlorinated polypropylene and polyethylenimine. Int. J. Hydrogen Energy 2017, 42, 23028–23034. [Google Scholar] [CrossRef]
  75. McKeown, N.B.; Budd, P.M. Polymers of Intrinsic Microporosity (PIMs): Organic Materials for Membrane Separations, Heterogeneous Catalysis and Hydrogen Storage. Chem. Soc. Rev. 2006, 35, 675–683. [Google Scholar] [CrossRef] [PubMed]
  76. Rochat, S.; Polak-Kraśna, K.; Tian, M.; Mays, T.J.; Bowen, C.R.; Burrows, A.D. Assessment of the Long-Term Stability of the Polymer of Intrinsic Microporosity PIM-1 for Hydrogen Storage Applications. State Art Mater. Hydrogen Energy 2019, 44, 332–337. [Google Scholar] [CrossRef]
  77. McKeown, N.B.; Gahnem, B.; Msayib, K.J.; Budd, P.M.; Tattershall, C.E.; Mahmood, K.; Tan, S.; Book, D.; Langmi, H.W.; Walton, A. Towards Polymer-based Hydrogen Storage Materials: Engineering Ultramicroporous Cavities within Polymers of Intrinsic Microporosity. Angew. Chem. Int. Ed. 2006, 45, 1804–1807. [Google Scholar] [CrossRef] [PubMed]
  78. McKeown, N.B.; Budd, P.M.; Book, D. Microporous Polymers as Potential Hydrogen Storage Materials. Macromol. Rapid Commun. 2007, 28, 995–1002. [Google Scholar] [CrossRef]
  79. Budd, P.M.; Butler, A.; Selbie, J.; Mahmood, K.; McKeown, N.B.; Ghanem, B.; Msayib, K.; Book, D.; Walton, A. The Potential of Organic Polymer-Based Hydrogen Storage Materials. Phys. Chem. Chem. Phys. 2007, 9, 1802–1808. [Google Scholar] [CrossRef]
  80. Ghanem, B.S.; Hashem, M.; Harris, K.D.; Msayib, K.J.; Xu, M.; Budd, P.M.; Chaukura, N.; Book, D.; Tedds, S.; Walton, A.; et al. Triptycene-Based Polymers of Intrinsic Microporosity: Organic Materials That Can Be Tailored for Gas Adsorption. Macromolecules 2010, 43, 5287–5294. [Google Scholar] [CrossRef]
  81. Bera, R.; Mondal, S.; Das, N. Triptycene Based Microporous Polymers (TMPs): Efficient Small Gas (H2 and CO2) Storage and High CO2/N2 Selectivity. Microporous Mesoporous Mater. 2018, 257, 253–261. [Google Scholar] [CrossRef]
  82. Makhseed, S.; Ibrahim, F.; Samuel, J. Phthalimide Based Polymers of Intrinsic Microporosity. Polymer 2012, 53, 2964–2972. [Google Scholar] [CrossRef]
  83. Weber, J.; Antonietti, M.; Tomas, A. Microporous Networks of High-Performance Polymers: Elastic Deformations and Gas Sorption Properties. Macromolecules 2008, 41, 2880–2885. [Google Scholar] [CrossRef]
  84. Ramimoghadam, D.; Gray, E.M.; Webb, C.J. Review of Polymers of Intrinsic Microporosity for Hydrogen Storage Applications. Int. J. Hydrogen Energy 2016, 41, 16944–16965. [Google Scholar] [CrossRef]
  85. Cooper, A.I. Conjugated Microporous Polymers. Adv. Mater. 2009, 21, 1291–1295. [Google Scholar] [CrossRef]
  86. Jiang, J.X.; Su, F.; Niu, H.; Wood, C.D.; Campbell, N.L.; Khimyak, Y.Z.; Cooper, A.I. Conjugated microporous poly(phenylene butadiynylene)s. Chem. Commun. 2008, 486–488. [Google Scholar] [CrossRef] [PubMed]
  87. Yuan, S.; Dorney, B.; White, D.; Kirklin, S.; Zapol, P.; Yu, L.; Liu, D.J. Microporous polyphenylenes with tunable pore size for hydrogen storage. Chem. Commun. 2010, 46, 4547–4549. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, Q.; Luo, M.; Wang, T.; Wang, J.X.; Zhou, D.; Han, Y.; Zhang, C.S.; Yan, C.G.; Han, B.H. Porous Organic Polymers Based on Propeller-like Hexaphenylbenzene Building Units. Macromolecules 2011, 44, 5573–5577. [Google Scholar] [CrossRef]
  89. Chen, Q.; Wang, J.X.; Yang, F.; Zhou, D.; Bian, N.; Zhang, X.J.; Yan, C.G.; Han, B.H. Tetraphenylethylene-Based Fluorescent Porous Organic Polymers: Preparation, Gas Sorption Properties and Photoluminescence Properties. J. Mater. Chem. 2011, 21, 13554–13560. [Google Scholar] [CrossRef]
  90. Suresh, V.M.; Bonakala, S.; Roy, S.; Balasubramanian, S.; Maji, T.K. Synthesis, characterization, and modeling of a functional conjugated microporous polymer: CO2 storage and light harvesting. J. Phys. Chem. C 2014, 118, 24369–24376. [Google Scholar] [CrossRef]
  91. Qiao, S.; Du, Z.; Yang, R. Design and Synthesis of Novel Carbazole–Spacer–Carbazole Type Conjugated Microporous Networks for Gas Storage and Separation. J. Mater. Chem. A 2014, 2, 1877–1885. [Google Scholar] [CrossRef]
  92. Bandyopadhyay, S.; Anil, A.G.; James, A.; Patra, A. Multifunctional Porous Organic Polymers: Tuning of Porosity, CO2, and H2 Storage and Visible-Light-Driven Photocatalysis. ACS Appl. Mater. Interfaces 2016, 8, 27669–27678. [Google Scholar] [CrossRef] [PubMed]
  93. Gu, C.; Bao, Y.; Huang, W.; Liu, D.; Yang, R. Four Simple Structure Carbazole-Based Conjugated Microporous Polymers with Different Soft Connected Chains. Macromol. Chem. Phys. 2016, 217, 748–756. [Google Scholar] [CrossRef]
  94. Rabbani, M.G.; Sekizkardes, A.K.; El-Kadri, O.M.; Kaafarani, B.R.; El-Kaderi, H.M. Pyrene-Directed Growth of Nanoporous Benzimidazole-Linked Nanofibers and Their Application to Selective CO2 Capture and Separation. J. Mater. Chem. 2012, 22, 25409–25417. [Google Scholar] [CrossRef]
  95. Bhunia, A.; Vasylyeva, V.; Janiak, C. From a Supramolecular Tetranitrile to a Porous Covalent Triazine-Based Framework with High Gas Uptake Capacities. Chem. Commun. 2013, 49, 3961–3963. [Google Scholar] [CrossRef]
  96. Kassab, R.M.; Jackson, K.T.; El-Kadri, O.M.; El-Kaderi, H.M. Nickel-Catalyzed Synthesis of Nanoporous Organic Frameworks and Their Potential Use in Gas Storage Applications. Res. Chem. Intermed. 2011, 37, 747–757. [Google Scholar] [CrossRef]
  97. Li, A.; Lu, R.F.; Wang, Y.; Wang, X.; Han, K.L.; Deng, W.Q. Lithium-doped Conjugated Microporous Polymers for Reversible Hydrogen Storage. Angew. Chem. 2010, 122, 3402–3405. [Google Scholar] [CrossRef]
  98. Xu, D.; Sun, L.; Li, G.; Shang, J.; Yang, R.X.; Deng, W.Q. Methyllithium-Doped Naphthyl-Containing Conjugated Microporous Polymer with Enhanced Hydrogen Storage Performance. Chem. Eur. J. 2016, 22, 7944–7949. [Google Scholar] [CrossRef] [PubMed]
  99. Reich, T.E.; Jackson, K.T.; Li, S.; Jena, P.; El-Kaderi, H.M. Synthesis and Characterization of Highly Porous Borazine-Linked Polymers and Their Performance in Hydrogen Storage Application. J. Mater. Chem. 2011, 21, 10629–10632. [Google Scholar] [CrossRef]
  100. Liao, Y.; Cheng, Z.; Zuo, W.; Thomas, A.; Fall, C.F. Nitrogen-Rich Conjugated Microporous Polymers: Facile Synthesis, Efficient Gas Storage, and Heterogeneous Catalysis. ACS Appl. Mater. 2017, 9, 38390–38400. [Google Scholar] [CrossRef] [PubMed]
  101. Xu, Y.; Li, Z.; Zhang, F.; Zhuang, X.; Zeng, Z.; Wei, J. New nitrogen-rich azo-bridged porphyrin-conjugated microporous networks for high performance of gas capture and storage. RSC Adv. 2016, 6, 30048–30055. [Google Scholar] [CrossRef]
  102. Sun, C.J.; Zhao, X.Q.; Wang, P.F.; Wang, H.; Han, B.H. Thiophene-based conjugated microporous polymers: Synthesis, characterization and efficient gas storage. Sci. China Chem. 2017, 60, 1067–1074. [Google Scholar] [CrossRef]
  103. Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J.M.; et al. Targeted Synthesis of a Porous Aromatic Framework with High Stability and Exceptionally High Surface Area. Angew. Chem. Int. Ed. 2009, 48, 9457–9460. [Google Scholar] [CrossRef]
  104. Lu, W.; Yuan, D.; Zhao, D.; Schilling, C.I.; Plietzsch, O.; Muller, T.; Bräse, S.; Guenther, J.; Blümel, J.; Krishna, R.; et al. Porous Polymer Networks: Synthesis, Porosity, and Applications in Gas Storage/Separation. Chem. Mater. 2010, 22, 5964–5972. [Google Scholar] [CrossRef]
  105. Yuan, D.; Lu, W.; Zhao, D.; Zhou, H.-C. Highly Stable Porous Polymer Networks with Exceptionally High Gas-uptake Capacities. Adv. Mater. 2011, 23, 3723–3725. [Google Scholar] [CrossRef] [PubMed]
  106. Wu, X.; Wang, R.; Yang, H.; Wang, W.; Cai, W.; Li, Q. Ultrahigh hydrogen storage capacity of novel porous aromatic frameworks. J. Mater. Chem. A 2015, 3, 10724–10729. [Google Scholar] [CrossRef]
  107. Huang, L.; Yang, X.; Cao, D. From inorganic to organic strategy to design porous aromatic frameworks for high-capacity gas storage. J. Phys. Chem. C 2015, 119, 3260–3267. [Google Scholar] [CrossRef]
  108. Konstas, K.; Taylor, J.W.; Thornton, A.W.; Doherty, C.M.; Lim, W.X.; Bastow, T.J.; Kennedy, D.F.; Wood, C.D.; Cox, B.J.; Hill, J.M.; et al. Lithiated Porous Aromatic Frameworks with Exceptional Gas Storage Capacity. Angew. Chem. 2012, 124, 6743–6746. [Google Scholar] [CrossRef]
  109. Ma, H.; Ren, H.; Zou, X.; Sun, F.; Yan, Z.; Cai, K.; Wang, D.; Zhu, G. Novel Lithium-Loaded Porous Aromatic Framework for Efficient CO2 and H2 Uptake. J. Mater. Chem. A 2013, 1, 752–758. [Google Scholar] [CrossRef]
  110. Ahmed, A.; Thornton, A.W.; Konstas, K.; Kannam, S.K.; Babarao, R.; Todd, B.D.; Hill, A.J.; Hill, M.R. Strategies toward Enhanced Low-Pressure Volumetric Hydrogen Storage in Nanoporous Cryoadsorbents. Langmuir 2013, 29, 15689–15697. [Google Scholar] [CrossRef]
  111. Wu, X.; Li, L.; Peng, L.; Wang, Y.; Cai, W. Effect of Coordinatively Unsaturated Metal Sites in Porous Aromatic Frameworks on Hydrogen Storage Capacity. Acta Phys. Chim. Sinca 2017, 34, 286–295. [Google Scholar]
  112. Demirocak, D.E.; Ram, M.K.; Srinivasan, S.S.; Goswami, D.Y.; Stefanakos, E.K. A Novel Nitrogen Rich Porous Aromatic Framework for Hydrogen and Carbon Dioxide Storage. J. Mater. Chem. A 2013, 1, 13800–13806. [Google Scholar] [CrossRef]
  113. Rochat, S.; Polak-Kraśna, K.; Tian, M.; Holyfield, L.T.; Mays, T.J.; Bowen, C.R.; Burrows, A.D. Hydrogen storage in polymer-based processable microporous composites. J. Mater. Chem. A 2017, 5, 18752–18761. [Google Scholar] [CrossRef] [Green Version]
  114. Lyu, H.; Zhang, Q.; Wang, Y.; Duan, J. Unified Meso-Pores and Dense Cu2+ Sites in Porous Coordination Polymers for Highly Efficient Gas Storage and Separation. Dalton Trans. 2018, 47, 4424–4427. [Google Scholar] [CrossRef]
  115. Kovalev, A.I.; Mart’yanova, E.S.; Khotina, I.A.; Klemenkova, Z.S.; Blinnikova, Z.K.; Volchkova, E.V.; Loginova, T.P.; Ponomarev, I.I. Polyphenylene Gels. Polym. Sci. Ser. B 2018, 60, 675–679. [Google Scholar] [CrossRef]
  116. Ahmed, S.D.; El-Hiti, A.G.; Yousif, E.; Hameed, S.A.; Abdalla, M. New Eco-Friendly Phosphorus Organic Polymers as Gas Storage Media. Polymers 2017, 9, 336. [Google Scholar] [CrossRef]
  117. Sing, K.S.W.; Williams, R.T. Physisorption Hysteresis Loops and the Characterization of Nanoporous Materials. Adsorpt. Sci. Technol. 2004, 22, 773–782. [Google Scholar] [CrossRef] [Green Version]
  118. Rouquerol, J.; Avnir, D.; Fairbridge, C.W.; Everett, D.H.; Haynes, J.H.; Pernicone, N.; Ramsay, J.D.F.; Sing, K.S.W.; Unger, K.K. Recommendations for the Characterization of Porous Solids. Pure Appl. Chem. 1994, 66, 1739–1758. [Google Scholar] [CrossRef]
  119. Horváth, G.; Kawazoe, K. Method for the Calculation of Effective Pore Size Distribution in Molecular Sieve Carbon. J. Chem. Eng. Jpn. 1983, 16, 470–475. [Google Scholar] [CrossRef]
  120. Cheng, L.S. Improved Horvath—Kawazoe Equations Including Spherical Pore Models for Calculating Micropore Size Distribution. Chem. Eng. Sci. 1994, 49, 2599–2609. [Google Scholar] [CrossRef]
  121. Dubinin, M.M. Fundamentals of the Theory of Adsorption in Micropores of Carbon Adsorbents: Characteristics of Their Adsorption Properties and Microporous Structures. Carbon 1989, 27, 457–467. [Google Scholar] [CrossRef]
  122. Dubinin, M.M. Chemistry and Physics of Carbon; Walker, L.P., Ed.; Marcel Dekker: New York, NY, USA, 1966; Volume 2, p. 51. [Google Scholar]
  123. Tarazona, P. Solid-Fluid Transition and Interfaces with Density Functional Approaches. Surf. Sci. 1995, 331, 989–994. [Google Scholar] [CrossRef]
  124. Zhechkov, L.; Heine, T.; Patchkovskii, S.; Seifert, G.; Duarte, H.A. An Efficient a Posteriori Treatment for Dispersion Interaction in Density-Functional-Based Tight Binding. J. Chem. Theory Comput. 2005, 1, 841–847. [Google Scholar] [CrossRef] [PubMed]
  125. Kolmogorov, A.N.; Crespi, V.H. Registry-Dependent Interlayer Potential for Graphitic Systems. Phys. Rev. B 2005, 71, 235415. [Google Scholar] [CrossRef]
  126. Breck, D. Zeolite Molecular Sieves: Structure, Chemistry, and Use; Krieger Pub Co.: Malabar, FL, USA, 1984. [Google Scholar]
  127. Lozano-Castelló, D.; Cazorla-Amoros, D.; Linares-Solano, A. Usefulness of CO2 Adsorption at 273 K for the Characterization of Porous Carbons. Carbon 2004, 42, 1233–1242. [Google Scholar] [CrossRef]
  128. Zhang, C.; Babonneau, F.; Bonhomme, C.; Laine, R.M.C.; Soles, L.; Hristov, H.A.; Yee, A.F. Highly Porous Polyhedral Silsesquioxane Polymers. Synthesis and Characterization. J. Am. Chem. Soc. 1998, 120, 8380–8391. [Google Scholar] [CrossRef]
  129. Zhang, R.; Phalen, R.N.; Cataquis, A.; Desta, M.; Kloesel, M. Study of Highly Porous Polymers for H2 Fuel Storage Using Positron Annihilation Lifetime Spectroscopy. Int. J. Hydrogen Energy 2015, 40, 8732–8741. [Google Scholar] [CrossRef]
  130. Jean, Y.C. Positron and Positronium Chemistry; World Scientific: River Edge, NJ, USA, 1990. [Google Scholar]
  131. Urban, C.; McCord, E.F.; Webster, O.W.; Abrams, L.; Long, H.W.; Gaede, H.; Tang, P.; Pines, A. 129Xe NMR Studies of Hyper-Cross-Linked Polyarylcarbinols: Rigid versus Flexible Structures. Chem. Matter 1995, 7, 1325–1332. [Google Scholar] [CrossRef]
  132. Zhang, C.; Liu, Y.; Li, B.; Tan, B.; Chen, C.F.; Xu, H.B.; Yang, X.L. Triptycene-Based Microporous Polymers: Synthesis and Their Gas Storage Properties. ACS Macro Lett. 2011, 1, 190–193. [Google Scholar] [CrossRef]
  133. Guo, J.H.; Zhang, H.; Tang, Y.; Cheng, X. Hydrogen Spillover Mechanism on Covalent Organic Frameworks as Investigated by Ab Initio Density Functional Calculation. Phys. Chem. Chem. Phys. 2013, 15, 2873–2881. [Google Scholar] [CrossRef]
  134. Lachawiec, A.J., Jr.; Yang, R.T. Isotope tracer study of hydrogen spillover on carbon-based adsorbents for hydrogen storage. Langmuir 2008, 24, 6159–6165. [Google Scholar] [CrossRef]
  135. Suri, M.; Dornfeld, M.; Ganz, E. Calculation of hydrogen storage capacity of metal-organic and covalent-organic frameworks by spillover. J. Chem. Phys. 2009, 131, 174703. [Google Scholar] [CrossRef] [PubMed]
  136. Yang, R.T.; Wang, Y. Catalyzed hydrogen spillover for hydrogen storage. J. Am. Chem. Soc. 2009, 131, 4224–4226. [Google Scholar] [CrossRef] [PubMed]
  137. Zhao, Y.; Gennett, T. Water-mediated cooperative migration of chemisorbed hydrogen on graphene. Phys. Rev. Lett. 2014, 112, 076101. [Google Scholar] [CrossRef] [PubMed]
  138. Ghimbeu, C.M.; Zlotea, C.; Gadiou, R.; Cuevas, F.; Leroy, E.; Latroche, M.; Vix-Guterl, C. Understanding the mechanism of hydrogen uptake at low pressure in carbon/palladium nanostructured composites. J. Mater. Chem. 2011, 21, 17765–17775. [Google Scholar] [CrossRef]
Figure 1. Theoretical simulation of adsorption potential for H2 ultra micropores with a few Å in radius: (a) schematic diagram of quantum sieving effect, (b) behavior of interaction potential, (c) adsorption potential versus pore size (Reproduced with permission from Ref. [32]. Copyright @ 2016 John Wiley and Sons).
Figure 1. Theoretical simulation of adsorption potential for H2 ultra micropores with a few Å in radius: (a) schematic diagram of quantum sieving effect, (b) behavior of interaction potential, (c) adsorption potential versus pore size (Reproduced with permission from Ref. [32]. Copyright @ 2016 John Wiley and Sons).
Polymers 11 00690 g001
Figure 2. H2 adsorption in pores with different sizes studied via in-situ neutron scattering (Reproduced with permission from [33]. Copyright @ 2011 ACS publications).
Figure 2. H2 adsorption in pores with different sizes studied via in-situ neutron scattering (Reproduced with permission from [33]. Copyright @ 2011 ACS publications).
Polymers 11 00690 g002
Figure 3. Schematic diagram of the different types of interactions between an H2 molecule and absorbent materials.
Figure 3. Schematic diagram of the different types of interactions between an H2 molecule and absorbent materials.
Polymers 11 00690 g003
Figure 4. Comparison of metal organic frameworks (MOFs), covalent organic frameworks (COFs), and active carbons in H2 storage at 77 K (Reproduced with permission from Ref. [50]. Copyright @ 2016 Elsevier).
Figure 4. Comparison of metal organic frameworks (MOFs), covalent organic frameworks (COFs), and active carbons in H2 storage at 77 K (Reproduced with permission from Ref. [50]. Copyright @ 2016 Elsevier).
Polymers 11 00690 g004
Figure 5. Basic synthetic route of hyper crosslinked polymers: polymers with aromatic side chains are linked using a high concentration of methylene dihalides in the presence of a Lewis acid catalyst, to create a high concentration of cross-links.
Figure 5. Basic synthetic route of hyper crosslinked polymers: polymers with aromatic side chains are linked using a high concentration of methylene dihalides in the presence of a Lewis acid catalyst, to create a high concentration of cross-links.
Polymers 11 00690 g005
Figure 6. Basic synthetic route of polymers of intrinsic microporosity.
Figure 6. Basic synthetic route of polymers of intrinsic microporosity.
Polymers 11 00690 g006
Figure 7. Basic structure of CTC-PIM, cyclotricatechylene-core polymer of intristic microporosity (a) and Trip(R)-PIM, a triptycene-core with variable R groups (b).
Figure 7. Basic structure of CTC-PIM, cyclotricatechylene-core polymer of intristic microporosity (a) and Trip(R)-PIM, a triptycene-core with variable R groups (b).
Polymers 11 00690 g007
Figure 8. Schematic synthetic strategy for polymers of intrinsic microporosity (PIMs) produced through condensation under mild reaction conditions: K2CO3, DMF, 60–120 °C (Reproduced with permission from [84]. Copyright @ 2016 Elsevier).
Figure 8. Schematic synthetic strategy for polymers of intrinsic microporosity (PIMs) produced through condensation under mild reaction conditions: K2CO3, DMF, 60–120 °C (Reproduced with permission from [84]. Copyright @ 2016 Elsevier).
Polymers 11 00690 g008aPolymers 11 00690 g008b
Figure 9. Basic synthetic route for conjugated microporous polymer-1 (CMP-1) and post-synthesis treatment. (Polymerization is promoted by Pd/Cu coupling under N2 atmosphere.)
Figure 9. Basic synthetic route for conjugated microporous polymer-1 (CMP-1) and post-synthesis treatment. (Polymerization is promoted by Pd/Cu coupling under N2 atmosphere.)
Polymers 11 00690 g009
Figure 10. CMPs with varied length linkers between branch points.
Figure 10. CMPs with varied length linkers between branch points.
Polymers 11 00690 g010
Figure 11. Basic synthetic route for porous aromatic framework-1 (PAF-1) (cross-coupling is promoted by Ni(0) catalyzed Yamamoto-type Ullman reaction under inert atmosphere.).
Figure 11. Basic synthetic route for porous aromatic framework-1 (PAF-1) (cross-coupling is promoted by Ni(0) catalyzed Yamamoto-type Ullman reaction under inert atmosphere.).
Polymers 11 00690 g011
Figure 12. Brief comparison between highly porous polymers and MOFs in H2 adsorption.
Figure 12. Brief comparison between highly porous polymers and MOFs in H2 adsorption.
Polymers 11 00690 g012
Figure 13. Schematic diagram of positron annihilation lifetime spectroscopy (PALS) technique: (a) setup and (b) mechanism.
Figure 13. Schematic diagram of positron annihilation lifetime spectroscopy (PALS) technique: (a) setup and (b) mechanism.
Polymers 11 00690 g013
Table 1. Major U.S. Department of Energy (DOE) Targets of Onboard Hydrogen Storage for Light Duty Vehicle.
Table 1. Major U.S. Department of Energy (DOE) Targets of Onboard Hydrogen Storage for Light Duty Vehicle.
Year20202025Ultimate
Target
Gravimetric capacity (wt %)4.5%5.5%6.5%
Volumetric capacity (g/L)304050
Cost ($/kg H2)333300266
Durability/Operability:
  • Operating temperature (°C)
  • Min/max delivery temperature (°C)
  • Operational cycles
  • Min/max delivery pressure (bar)
  • Onboard efficiency
−40/60
−40/85
1500
5/12
90%
−40/60
−40/85
1500
5/12
90%
−40/60
−40/85
1500
5/12
90%
Charging/Discharging rate:
  • System fill time (min)
  • Min full flow rate ((g/s)/kW)
  • Average flow rate ((g/s)/kW)
  • Start time to full flow @ 20 °C (s)
  • Start time to full flow @ −20 °C (s)
  • Transient response at operating temperature 10%–90% and 90%–0% (based on full flow rate) (s)
3–5
0.02
0.004
5
15
0.75
3–5
0.02
0.004
5
15
0.75
3–5
0.02
0.004
5
15
0.75
Table 2. Dependence on distance and magnitude of different type of interactions between an H2 and adsorbent molecules (Modified from Ref. [35]).
Table 2. Dependence on distance and magnitude of different type of interactions between an H2 and adsorbent molecules (Modified from Ref. [35]).
Interaction type (Material – H2)Energy dependenceTypical values (kJ/mol)
Charge – H2 quadropole∝ 1/r 3~ 3.5
Charge – induced H2 dipole∝ 1/r 4~ 6.8
Dipole – induced H2 dipole∝ 1/r 5~ 0.6
van der Waals∝ 1/r 6~ 5 – 6
Orbital interaction<vdW radii~ 20 – 160
Table 3. Hyper crosslinked polymers (HCPs) and their H2 uptake.
Table 3. Hyper crosslinked polymers (HCPs) and their H2 uptake.
MonomerCrosslinkerSSA (m2/g)H2 UptakeRef.
VBC-DVBVBC1300–19003 wt % 77 K/15 bar,
1.5 wt % 77 K/1 bar
[52,53]
VCBDVB15001.59 wt % 77 K/1.13 bar[65]
Carbazole bromophenyl- methanol1,10-Phenantroline10002.39 wt % 77 K/1 bar[66]
CarbazolesFDA1000–18001.94 wt % 77 K/1 bar[67]
CarbazolesDCX600–19000.9–17.7 wt % 77 K/1 bar[59]
PolyanilinesCH2I2~5002.2 wt % 77 K/30 bar[57]
PolypyrrolesCH2I2~20–7000.6–1.6 wt % 77 K/4 bar[58]
BismaleimidesDVB8410.82 wt % 77 K/1 bar[68]
TPBCl17831.91 wt % 77 K/1 bar[69]
TPBFDA10591.58 wt % 77 K/1.13 bar[70]
BenzeneFDA13911.45 wt % 77 K/1.13 bar[70]
FluoreneBCMBP17001.63 wt % 77 K/1 bar[71]
TPEFDA19801.76 wt % 77 K/1 bar[72]
VBC-DVB: vinylbenzyl chloride-divinylbenzene; VBC: vinylbenzyl chloride; DCX: dichloroxylene; DVB: divinylbenzene; FDA: formaldehyde dimethyl acetal; TPB: 1,3,5-triphenylbenzene; BCMBP: 4,4’-bis(chloromethyl)biphenyl; TPE: tetraphenylethylene.
Table 4. Comparison of specific surface area, pore size, and H2 uptake in porous organic polymers (POPs) (Adapted from [87])
Table 4. Comparison of specific surface area, pore size, and H2 uptake in porous organic polymers (POPs) (Adapted from [87])
CMPsSBET (m2/g)Pore diamter (nm)Micropore volume (cm3/g)H2 uptake
POP-110310.770.3782.78 wt % 77K/60bar
POP-210130.740.3412.71 wt % 77K/60bar
POP-312460.880.4483.07 wt % 77K/60bar
POP-410330.810.4022.35 wt % 77K/60bar

Share and Cite

MDPI and ACS Style

Cousins, K.; Zhang, R. Highly Porous Organic Polymers for Hydrogen Fuel Storage. Polymers 2019, 11, 690. https://doi.org/10.3390/polym11040690

AMA Style

Cousins K, Zhang R. Highly Porous Organic Polymers for Hydrogen Fuel Storage. Polymers. 2019; 11(4):690. https://doi.org/10.3390/polym11040690

Chicago/Turabian Style

Cousins, Kimberley, and Renwu Zhang. 2019. "Highly Porous Organic Polymers for Hydrogen Fuel Storage" Polymers 11, no. 4: 690. https://doi.org/10.3390/polym11040690

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop