Hydrogen storage

Utility scale underground liquid hydrogen storage

Methods of hydrogen storage for subsequent use span many approaches including high pressures, cryogenics, and chemical compounds that reversibly release H2 upon heating. Underground hydrogen storage is useful to provide grid energy storage for intermittent energy sources, like wind power, as well as providing fuel for transportation, particularly for ships and airplanes.

Most research into hydrogen storage is focused on storing hydrogen as a lightweight, compact energy carrier for mobile applications.

Liquid hydrogen or slush hydrogen may be used, as in the Space Shuttle. However liquid hydrogen requires cryogenic storage and boils around 20.268 K (−252.882 °C or −423.188 °F). Hence, its liquefaction imposes a large energy loss (as energy is needed to cool it down to that temperature). The tanks must also be well insulated to prevent boil off but adding insulation increases cost. Liquid hydrogen has less energy density by volume than hydrocarbon fuels such as gasoline by approximately a factor of four. This highlights the density problem for pure hydrogen: there is actually about 64% more hydrogen in a liter of gasoline (116 grams hydrogen) than there is in a liter of pure liquid hydrogen (71 grams hydrogen). The carbon in the gasoline also contributes to the energy of combustion.

Compressed hydrogen, by comparison, is stored quite differently. Hydrogen gas has good energy density by weight, but poor energy density by volume versus hydrocarbons, hence it requires a larger tank to store. A large hydrogen tank will be heavier than the small hydrocarbon tank used to store the same amount of energy, all other factors remaining equal. Increasing gas pressure would improve the energy density by volume, making for smaller, but not lighter container tanks (see hydrogen tank). Compressed hydrogen costs 2.1% of the energy content[1] to power the compressor. Higher compression without energy recovery will mean more energy lost to the compression step. Compressed hydrogen storage can exhibit very low permeation.[2]

Automotive Onboard hydrogen storage

Targets were set by the FreedomCAR Partnership in January 2002 between the United States Council for Automotive Research (USCAR) and U.S. DOE (Targets assume a 5-kg H2 storage system). The 2005 targets were not reached in 2005.[3] The targets were revised in 2009 to reflect new data on system efficiencies obtained from fleets of test cars.[4] The ultimate goal for volumetric storage is still above the theoretical density of liquid hydrogen.[5]

It is important to note that these targets are for the hydrogen storage system, not the hydrogen storage material. System densities are often around half those of the working material, thus while a material may store 6 wt% H2, a working system using that material may only achieve 3 wt% when the weight of tanks, temperature and pressure control equipment, etc., is considered.

In 2010, only two storage technologies were identified as having the potential to meet DOE targets: MOF-177 exceeds 2010 target for volumetric capacity, while cryo-compressed H2 exceeds more restrictive 2015 targets for both gravimetric and volumetric capacity (see slide 6 in [6]).

Established technologies

net storage density of hydrogen

Compressed hydrogen

Compressed hydrogen is a storage form where hydrogen gas is kept under pressures to increase the storage density. Compressed hydrogen in hydrogen tanks at 350 bar (5,000 psi) and 700 bar (10,000 psi) is used for hydrogen tank systems in vehicles, based on type IV carbon-composite technology.[7] Car manufacturers have been developing this solution, such as Honda[8] or Nissan.[9]

Liquid hydrogen

BMW has been working on liquid hydrogen tanks for cars, producing for example the BMW Hydrogen 7.

Proposals and research

Hydrogen storage technologies can be divided into physical storage, where hydrogen molecules are stored (including pure hydrogen storage via compression and liquefaction), and chemical storage, where hydrides are stored.

Chemical storage

Chemical storage could offer high storage performance due to the strong binding of hydrogen and the high storage densities. However, the regeneration of storage material is still an issue. A large number of chemical storage systems are under investigation, which involve hydrolysis reactions, hydrogenation/dehydrogenation reactions, ammonia borane and other boron hydrides, ammonia, and alane etc.[10] Storage in hydrocarbons may also be successful in overcoming the issue with low density. For example, supercritical hydrogen at 30 °C and 500 bar only has a density of 15.0 mol/L while methanol has a density of 49.5 mol H2/L methanol and saturated dimethyl ether at 30 °C and 7 bar has a density of 42.1 mol H2/L dimethyl ether. These liquids would use much smaller, cheaper, safer storage tanks.

Metal hydrides
Metal hydride hydrogen storage

Metal hydrides, such as MgH2, NaAlH4, LiAlH4, LiH, LaNi5H6, TiFeH2 and palladium hydride, with varying degrees of efficiency, can be used as a storage medium for hydrogen, often reversibly.[11] Some are easy-to-fuel liquids at ambient temperature and pressure, others are solids which could be turned into pellets. These materials have good energy density by volume, although their energy density by weight is often worse than the leading hydrocarbon fuels.

Most metal hydrides bind with hydrogen very strongly. As a result, high temperatures around 120 °C (248 °F) – 200 °C (392 °F) are required to release their hydrogen content. This energy cost can be reduced by using alloys which consists of a strong hydride former and a weak one such as in LiNH2, LiBH4 and NaBH4.[12] These are able to form weaker bonds, thereby requiring less input to release stored hydrogen. However, if the interaction is too weak, the pressure needed for rehydriding is high, thereby eliminating any energy savings. The target for onboard hydrogen fuel systems is roughly <100 °C for release and <700 bar for recharge (20–60 kJ/mol H2).[13]

An alternative method for reducing dissociation temperatures is doping with activators. This has been successfully used for aluminium hydride but its complex synthesis makes it undesirable for most applications as it is not easily recharged with hydrogen.[14]

Currently the only hydrides which are capable of achieving the 9 wt. % gravimetric goal for 2015 (see chart above) are limited to lithium, boron and aluminium based compounds; at least one of the first-row elements or Al must be added. Research is being done to determine new compounds which can be used to meet these requirements.

Proposed hydrides for use in a hydrogen economy include simple hydrides of magnesium[15] or transition metals and complex metal hydrides, typically containing sodium, lithium, or calcium and aluminium or boron. Hydrides chosen for storage applications provide low reactivity (high safety) and high hydrogen storage densities. Leading candidates are lithium hydride, sodium borohydride, lithium aluminium hydride and ammonia borane. A French company McPhy Energy is developing the first industrial product, based on magnesium hydride, already sold to some major clients such as Iwatani and ENEL.

New Scientist reported that Arizona State University is investigating using a borohydride solution to store hydrogen, which is released when the solution flows over a catalyst made of ruthenium.[16] Researchers at University of Pittsburgh and Georgia Tech performed extensive benchmarking simulations on mixtures of several light metal hydrides to predict possible reaction thermodynamics for hydrogen storage.[17][18][19]

Non-metal hydrides

The Italian catalyst manufacturer Acta has proposed using hydrazine as an alternative to hydrogen in fuel cells. As the hydrazine fuel is liquid at room temperature, it can be handled and stored more easily than hydrogen. By storing it in a tank full of a double-bonded carbon-oxygen carbonyl, it reacts and forms a safe solid called hydrazone. By then flushing the tank with warm water, the liquid hydrazine hydrate is released. Hydrazine breaks down in the cell to form nitrogen and hydrogen which bonds with oxygen, releasing water.[20]

Carbohydrates

Carbohydrates (polymeric C6H10O5) releases H2 in a bioreformer mediated by the enzyme cocktail—cell-free synthetic pathway biotransformation. Carbohydrate provides high hydrogen storage densities as a liquid with mild pressurization and cryogenic constraints: It can also be stored as a solid powder. Carbohydrate is the most abundant renewable bioresource in the world.

In May 2007 biochemical engineers from the Virginia Polytechnic Institute and State University and biologists and chemists from the Oak Ridge National Laboratory announced a method of producing high-yield pure hydrogen from starch and water.[21] In 2009, they demonstrated to produce nearly 12 moles of hydrogen per glucose unit from cellulosic materials and water.[22] Thanks to complete conversion and modest reaction conditions, they propose to use carbohydrate as a high energy density hydrogen carrier with a density of 14.8 wt%.[23]

Synthesized hydrocarbons

An alternative to hydrides is to use regular hydrocarbon fuels as the hydrogen carrier. Then a small hydrogen reformer would extract the hydrogen as needed by the fuel cell. However, these reformers are slow to react to changes in demand and add a large incremental cost to the vehicle powertrain.

Direct methanol fuel cells do not require a reformer, but provide a lower energy density compared to conventional fuel cells, although this could be counterbalanced with the much better energy densities of ethanol and methanol over hydrogen. Alcohol fuel is a renewable resource.

Solid-oxide fuel cells can operate on light hydrocarbons such as propane and methane without a reformer, or can run on higher hydrocarbons with only partial reforming, but the high temperature and slow startup time of these fuel cells are problematic for automotive applications.

Liquid organic hydrogen carriers (LOHC)

Unsaturated organic compounds can store huge amounts of hydrogen. These Liquid Organic Hydrogen Carriers (LOHC) are hydrogenated for storage and dehydrogenated again when the energy/hydrogen is needed. Research on LOHC was concentrated on cycloalkanes at an early stage, with its relatively high hydrogen capacity (6-8 wt %) and production of COx-free hydrogen.[24] Heterocyclic aromatic compounds (or N-Heterocycles) are also appropriate for this task. A compound that stands in the focus of the current LOHC research is N-ethylcarbazole (NEC)[25] but many others do exist.[26] More recently dibenzyltoluene, which is already industrially used as a heat transfer fluid in industry, was identified as potential LOHC. With a wide liquid range between -39 °C (melting point) and 390 °C (boiling point) and a hydrogen storage density of 6.2 wt.-% dibenzyltoluene is ideally suited as LOHC material.[27] More recently, formic acid (FA) has been suggested as a promising hydrogen storage material with a 4.4wt% hydrogen capacity.[28]

Using LOHCs relatively high gravimetric storage densities can be reached (about 6 wt-%) and the overall energy efficiency is higher than for other chemical storage options such as producing methane from the hydrogen.[29]

Cycloalkanes reported as LOHC include cyclohexane, methyl-cyclohexane and decalin. The dehydrogenation of cycloalkanes is highly endothermic (63-69 kJ/mol H2), which means this process requires high temperature.[24] Dehydrogenation of decalin is the most thermodynamically favored among the three cycloalkanes, and methyl-cyclohexane is second because of the presence of the methyl group.[30] Research on catalyst development for dehydrogenation of cycloalkanes has been carried out for decades. Nickel (Ni), Molybdenum (Mo) and Platinum (Pt) based catalysts are highly investigated for dehydrogenation. However, coking is still a big challenge for catalyst’s long-term stability.[31][32]

Both hydrogenation and dehydrogenation of LOHCs requires catalysts.[24] It was demonstrated that replacing hydrocarbons by hetero-atoms, like N, O etc. improves reversible de/hydrogenation properties. The temperature required for hydrogenation and dehydrogenation of drops significantly with increasing numbers of heteoatoms.[33] Among all the N-heterocycles, the saturated-unsaturated pair of dodecahydro-N-ethylcarbazole (12H-NEC) and NEC has been considered as a promising candidate for hydrogen storage with a fairly large hydrogen content (5.8wt%).[34] The figure on the top right shows dehydrogenation and hydrogenation of the 12H-NEC and NEC pair. The standard catalyst for NEC to 12H-NEC is Ru and Rh based. The selectivity of hydrogenation can reach 97% at 7 MPa and 130 °C-150 °C.[24] Although N-Heterocyles can optimize the unfavorable thermodynamic properties of cycloalkanes, a lot of issues remain unsolved, such as high cost, high toxicity and kinetic barriers etc.[24]

In 2006 researchers of EPFL, Switzerland, reported the use of formic acid as a hydrogen storage material.[35] Carbon monoxide free hydrogen has been generated in a very wide pressure range (1–600 bar). A homogeneous catalytic system based on water-soluble ruthenium catalysts selectively decompose HCOOH into H2 and CO2 in aqueous solution.[36] This catalytic system overcomes the limitations of other catalysts (e.g. poor stability, limited catalytic lifetimes, formation of CO) for the decomposition of formic acid making it a viable hydrogen storage material.[37] And the co-product of this decomposition, carbon dioxide, can be used as hydrogen vector by hydrogenating it back to formic acid in a second step. The catalytic hydrogenation of CO2 has long been studied and efficient procedures have been developed.[38][39] Formic acid contains 53 g L−1 hydrogen at room temperature and atmospheric pressure. By weight, pure formic acid stores 4.3 wt% hydrogen. Pure formic acid is a liquid with a flash point 69 °C (cf. gasoline −40 °C, ethanol 13 °C). 85% formic acid is not flammable.

Ammonia
Further information: Ammonia production

Ammonia (NH3) releases H2 in an appropriate catalytic reformer. Ammonia provides high hydrogen storage densities as a liquid with mild pressurization and cryogenic constraints: It can also be stored as a liquid at room temperature and pressure when mixed with water. Ammonia is the second most commonly produced chemical in the world and a large infrastructure for making, transporting, and distributing ammonia exists. Ammonia can be reformed to produce hydrogen with no harmful waste, or can mix with existing fuels and under the right conditions burn efficiently. Since there is no carbon in ammonia, no carbon by-products are produced; thereby making this possibility a "carbon neutral" option for the future. Pure ammonia burns poorly at the atmospheric pressures found in natural gas fired water heaters and stoves. Under compression in an automobile engine it is a suitable fuel for slightly modified gasoline engines. Ammonia is a toxic gas at normal temperature and pressure and has a potent odor.[40]

In September 2005 chemists from the Technical University of Denmark announced a method of storing hydrogen in the form of ammonia saturated into a salt tablet. They claim it will be an inexpensive and safe storage method.[41]

Amine borane complexes
Main article: Amine borane complex

Prior to 1980, several compounds were investigated for hydrogen storage including complex borohydrides, or aluminohydrides, and ammonium salts. These hydrides have an upper theoretical hydrogen yield limited to about 8.5% by weight. Amongst the compounds that contain only B, N, and H (both positive and negative ions), representative examples include: amine boranes, boron hydride ammoniates, hydrazine-borane complexes, and ammonium octahydrotriborates or tetrahydroborates. Of these, amine boranes (and especially ammonia borane) have been extensively investigated as hydrogen carriers. During the 1970s and 1980s, the U.S. Army and Navy funded efforts aimed at developing hydrogen/deuterium gas-generating compounds for use in the HF/DF and HCl chemical lasers, and gas dynamic lasers. Earlier hydrogen gas-generating formulations used amine boranes and their derivatives. Ignition of the amine borane(s) forms boron nitride (BN) and hydrogen gas. In addition to ammonia borane (H3BNH3), other gas-generators include diborane diammoniate, H2B(NH3)2BH4.

Imidazolium ionic liquids

In 2007 Dupont and others reported hydrogen-storage materials based on imidazolium ionic liquids. Simple alkyl(aryl)-3-methylimidazolium N-bis(trifluoromethanesulfonyl)imidate salts that possess very low vapour pressure, high density, and thermal stability and are not inflammable can add reversibly 6–12 hydrogen atoms in the presence of classical Pd/C or Ir0 nanoparticle catalysts and can be used as alternative materials for on-board hydrogen-storage devices. These salts can hold up to 30 g L−1 of hydrogen at atmospheric pressure.[42]

Phosphonium borate

In 2006 researchers of University of Windsor reported on reversible hydrogen storage in a non-metal phosphonium borate frustrated Lewis pair:[43][44][45]

The phosphino-borane on the left accepts one equivalent of hydrogen at one atmosphere and 25 °C and expels it again by heating to 100 °C. The storage capacity is 0.25 wt% still rather below the 6 to 9 wt% required for practical use.

Carbonite substances

Research has proven that graphene can store hydrogen efficiently. After taking up hydrogen, the substance becomes graphane. After tests, conducted by dr André Geim at the University of Manchester, it was shown that not only can graphene store hydrogen easily, it can also release the hydrogen again, after heating to 450 °C.[46][47]

Metal-organic frameworks

Metal-organic frameworks represent another class of synthetic porous materials that store hydrogen and energy at the molecular level. MOFs are highly crystalline inorganic-organic hybrid structures that contain metal clusters or ions (secondary building units) as nodes and organic ligands as linkers. When guest molecules (solvent) occupying the pores are removed during solvent exchange and heating under vacuum, porous structure of MOFs can be achieved without destabilizing the frame and hydrogen molecules will be adsorbed onto the surface of the pores by physisorption. Compared to traditional zeolites and porous carbon materials, MOFs have very high number of pores and surface area which allow higher hydrogen uptake in a given volume. Thus, research interests on hydrogen storage in MOFs have been growing since 2003 when the first MOF-based hydrogen storage was introduced. Since there are infinite geometric and chemical variations of MOFs based on different combinations of SBUs and linkers, many researches explore what combination will provide the maximum hydrogen uptake by varying materials of metal ions and linkers.

In 2006, chemists at UCLA and the University of Michigan have achieved hydrogen storage concentrations of up to 7.5 wt% in MOF-74 at a low temperature of 77 K.[48][49] In 2009, researchers at University of Nottingham reached 10 wt% at 77 bar (1,117 psi) and 77 K with MOF NOTT-112.[50] Most articles about hydrogen storage in MOFs report hydrogen uptake capacity at a temperature of 77K and a pressure of 1 bar because such condition is commonly available and the binding energy between hydrogen and MOF is large compare to the thermal vibration energy which will allow high hydrogen uptake capacity. Varying several factors such as surface area, pore size, catenation, ligand structure, spillover, and sample purity can result different amount of hydrogen uptake in MOFs.

Encapsulation

Cella Energy technology is based around the encapsulation of hydrogen gas and nano-structuring of chemical hydrides in small plastic balls, at room temperature and pressure.[51]

Physical storage

In this case hydrogen remains in physical forms, i.e., as gas, supercritical fluid, adsorbate, or molecular inclusions. Theoretical limitations and experimental results are considered [52] concerning the volumetric and gravimetric capacity of glass microvessels, microporous, and nanoporous media, as well as safety and refilling-time demands.

Cryo-compressed

Cryo-compressed storage of hydrogen is the only technology that meets 2015 DOE targets for volumetric and gravimetric efficiency (see "CcH2" on slide 6 in [6]).

Furthermore, another study has shown that cryo-compressed exhibits interesting cost advantages: ownership cost (price per mile) and storage system cost (price per vehicle) are actually the lowest when compared to any other technology (see third row in slide 13 of [53]). For example, a cryo-compressed hydrogen system would cost $0.12 per mile (including cost of fuel and every associated other cost), while conventional gasoline vehicles cost between $0.05 and $0.07 per mile.

Like liquid storage, cryo-compressed uses cold hydrogen (20.3 K and slightly above) in order to reach a high energy density. However, the main difference is that, when the hydrogen would warm-up due to heat transfer with the environment ("boil off"), the tank is allowed to go to pressures much higher (up to 350 bars versus a couple of bars for liquid storage). As a consequence, it takes more time before the hydrogen has to vent, and in most driving situations, enough hydrogen is used by the car to keep the pressure well below the venting limit.

Consequently, it has been demonstrated that a high driving range could be achieved with a cryo-compressed tank : more than 650 miles (1,050 km) were driven with a full tank mounted on an hydrogen-fueled engine of Toyota Prius.[54] Research is still on its way in order to study and demonstrate the full potential of the technology.[55]

As of 2010, the BMW Group has started a thorough component and system level validation of cryo-compressed vehicle storage on its way to a commercial product.[56]

Carbon nanotubes
Carbon nanotubes

Hydrogen carriers based on nanostructured carbon (such as carbon buckyballs and nanotubes) have been proposed. However, since Hydrogen usually amounts up to ~3.0-7.0 wt.% at 77K which is far from the value set by US department of Energy (6wt.% at nearly ambient conditions), it makes carbon materials poor candidates for hydrogen storage.

Clathrate hydrates

H2 caged in a clathrate hydrate was first reported in 2002, but requires very high pressures to be stable. In 2004, researchers from Delft University of Technology and Colorado School of Mines showed solid H2-containing hydrates could be formed at ambient temperature and 10s of bar by adding small amounts of promoting substances such as THF.[57] These clathrates have a theoretical maximum hydrogen densities of around 5 wt% and 40 kg/m3.

Glass capillary arrays

A team of Russian, Israeli and German scientists have collaboratively developed an innovative technology based on glass capillary arrays for the safe infusion, storage and controlled release of hydrogen in mobile applications.[58][59] The C.En technology has achieved the United States Department of Energy (DOE) 2010 targets for on-board hydrogen storage systems.[60] DOE 2015 targets can be achieved using flexible glass capillaries and cryo-compressed method of hydrogen storage [61]

Glass microspheres

Hollow glass microspheres (HGM) can be utilized for controlled storage and release of hydrogen.[62][63]

Stationary hydrogen storage

Unlike mobile applications, hydrogen density is not a huge problem for stationary applications. As for mobile applications, stationary applications can use established technology:

Underground hydrogen storage

Underground hydrogen storage is the practice of hydrogen storage in underground caverns, salt domes and depleted oil and gas fields. Large quantities of gaseous hydrogen have been stored in underground caverns by ICI for many years without any difficulties.[65] The storage of large quantities of liquid hydrogen underground can function as grid energy storage. The round-trip efficiency is approximately 40% (vs. 75-80% for pumped-hydro (PHES)), and the cost is slightly higher than pumped hydro.[66] The European project Hyunder[67] indicated in 2013 that for the storage of wind and solar energy an additional 85 caverns are required as it can't be covered by PHES and CAES systems.[68]

Power to gas

Power to gas is a technology which converts electrical power to a gas fuel. There are two methods: the first is to use the electricity for water splitting and inject the resulting hydrogen into the natural gas grid; the second, less efficient method is used to convert carbon dioxide and hydrogen to methane, (see natural gas) using electrolysis and the Sabatier reaction. The excess power or off peak power generated by wind generators or solar arrays is then used for load balancing in the energy grid. Using the existing natural gas system for hydrogen Fuel cell maker Hydrogenics and natural gas distributor Enbridge have teamed up to develop such a power to gas system in Canada.[69]

Pipeline storage of hydrogen where a natural gas network is used for the storage of hydrogen. Before switching to natural gas, the German gas networks were operated using towngas, which for the most part (60-65%) consisted of hydrogen. The storage capacity of the German natural gas network is more than 200,000 GW·h which is enough for several months of energy requirement. By comparison, the capacity of all German pumped storage power plants amounts to only about 40 GW·h. The transport of energy through a gas network is done with much less loss (<0.1%) than in a power network (8%). The use of the existing natural gas pipelines for hydrogen was studied by NaturalHy[70]

See also

References

  1. Energy technology analysis. International Energy Agency (2005) p. 70
  2. Modeling of dispersion following hydrogen permeation for safety engineering and risk assessment. (PDF) . II International Conference "Hydrogen Storage Technologies" Moscow, Russia, 28–29 October 2009. Retrieved on 2012-01-08.
  3. Hydrogen Storage Technologies Roadmap. uscar.org. November 2005
  4. Yang, Jun; Sudik, A; Wolverton, C; Siegel, D.J. (2010). "High capacity hydrogen storage materials: attributes for automotive applications and techniques for materials discovery". Chem. Soc. Rev. 39 (2): 656–675. doi:10.1039/b802882f. PMID 20111786.
  5. FCT Hydrogen Storage: Current Technology. eere.energy.gov (2011-08-26). Retrieved on 2012-01-08.
  6. 1 2 R. K. Ahluwalia, T. Q. Hua, J. K. Peng and R. Kumar System Level Analysis of Hydrogen Storage Options. 2010 DOE Hydrogen Program Review, Washington, DC, June 8–11, 2010
  7. Eberle, Ulrich; Mueller, Bernd; von Helmolt, Rittmar. "Fuel cell electric vehicles and hydrogen infrastructure: status 2012". Energy & Environmental Science. Retrieved 2014-12-19.
  8. Honda Worldwide | FCX Clarity. World.honda.com. Retrieved on 2012-01-08.
  9. NISSAN | TECHNOLOGICAL DEVELOPMENT ACTIVITIES | Overview | X-TRAIL FCV '03 model. Nissan-global.com. Retrieved on 2012-01-08.
  10. Sunita, Satyapal (2007). "The U.S. Department of Energy's National Hydrogen Storage Project: Progress towards meeting hydrogen-powered vehicle requirements". Catalysis Today. 120: 246–256. doi:10.1016/j.cattod.2006.09.022.
  11. DOE Metal hydrides. eere.energy.gov (2008-12-19). Retrieved on 2012-01-08.
  12. Christian, Meganne; Aguey-Zinsou, Kondo François (2012). "Core–Shell Strategy Leading to High Reversible Hydrogen Storage Capacity for NaBH4". ACS Nano. American Chemical Society. 6: 7739–7751. doi:10.1021/nn3030018. Retrieved 20 August 2012.
  13. EU Hydrogen Storage. (PDF) . Retrieved on 2012-01-08.
  14. Graetz, J.; Reilly, J.; Sandrock, G.; Johnson, J.; Zhou, W. M.; Wegrzyn, J. (2006). "Aluminum Hydride, A1H3, As a Hydrogen Storage Compound". doi:10.2172/899889.
  15. CNRS Institut Neel H2 Storage. Neel.cnrs.fr. Retrieved on 2012-01-08.
  16. "New type of hydrogen fuel cell powers up". newscientist. Retrieved 2006-09-16.
  17. Ki Chul Kim; Anant D. Kulkarni; J. Karl Johnson; David S. Sholl (2011). "Examining the robustness of first-principles calculations for metal hydride reaction thermodynamics by detection of metastable reaction pathways". Phys. Chem. Chem. Phys. 13: 21520. doi:10.1039/C1CP22489A.
  18. Ki Chul Kim; Anant D. Kulkarni; J. Karl Johnson; David S. Sholl (2011). "Large-scale screening of promising metal hydrides for hydrogen storage system from first-principles calculations based on equilibrium reaction thermodynamics". Phys. Chem. Chem. Phys. 13: 7218. doi:10.1039/c0cp02950e.
  19. Anant D. Kulkarni; Lin-Lin Wang; Duane D. Johnson; David S. Sholl; J. Karl Johnson (2010). "First-Principles Characterization of Amorphous Phases of MB12H12, M = Mg, Ca". J. Phys. Chem. C. 114: 14601–14605. doi:10.1021/jp101326g.
  20. "Liquid asset". The Engineer. 2008-01-15. Retrieved 2015-01-09.
  21. Zhang, Y.-H. Percival; Evans, Barbara R.; Mielenz, Jonathan R.; Hopkins, Robert C.; Adams, Michael W.W. (2007). Melis, Anastasios, ed. "High-Yield Hydrogen Production from Starch and Water by a Synthetic Enzymatic Pathway". PLoS ONE. 2 (5): e456. doi:10.1371/journal.pone.0000456. PMC 1866174Freely accessible. PMID 17520015.
  22. Ye, Xinhao; Wang, Yiran; Hopkins, Robert C.; Adams, Michael W. W.; Evans, Barbara R.; Mielenz, Jonathan R.; Zhang, Y.-H. Percival (2009). "Spontaneous High-Yield Production of Hydrogen from Cellulosic Materials and Water Catalyzed by Enzyme Cocktails". ChemSusChem. 2 (2): 149–52. doi:10.1002/cssc.200900017. PMID 19185036.
  23. Zhang, Y.-H. Percival (2009). "A sweet out-of-the-box solution to the hydrogen economy: is the sugar-powered car science fiction?". Energy & Environmental Science. 2 (3): 272. doi:10.1039/b818694d.
  24. 1 2 3 4 5 He, Teng; Pei, Qijun; Chen, Ping (2015-09-01). "Liquid organic hydrogen carriers". Journal of Energy Chemistry. 24 (5): 587–594. doi:10.1016/j.jechem.2015.08.007.
  25. D. Teichmann, W. Arlt, P. Wasserscheid, R. Freymann, „A future energy supply based on Liquid Organic Hydrogen Carriers (LOHC)", Energy Environ. Sci., 2011,4, 2767–2773
  26. US patent 7351395, "Hydrogen storage by reversible hydrogenation of pi-conjugated substrates"
  27. N. Brückner, K. Obesser, Dr. A. Bösmann, D. Teichmann, Prof. W. Arlt, Dr. J. Dungs, Prof. P. Wasserscheid, „Evaluation of Industrially Applied Heat-Transfer Fluids as Liquid Organic Hydrogen Carrier Systems", ChemSusChem, 2013
  28. Grasemann, Martin; Laurenczy, Gábor. "Formic acid as a hydrogen source – recent developments and future trends". pubs.rsc.org. doi:10.1039/C2EE21928J. Retrieved 2015-11-04.
  29. B. Müller, K. Müller, D. Teichmann, W. Arlt, "Energy Storage by CO2 Methanization and Energy Carrying Compounds: A Thermodynamic Comparison", Chemie Ingenieur Technik, 2011,11, 2002–2013 (written in German)
  30. Wang, Bo; Goodman, D. Wayne; Froment, Gilbert F. (2008-01-25). "Kinetic modeling of pure hydrogen production from decalin". Journal of Catalysis. 253 (2): 229–238. doi:10.1016/j.jcat.2007.11.012.
  31. Kariya, Nobuko; Fukuoka, Atsushi; Ichikawa, Masaru (2002-07-10). "Efficient evolution of hydrogen from liquid cycloalkanes over Pt-containing catalysts supported on active carbons under "wet–dry multiphase conditions"". Applied Catalysis A: General. 233 (1–2): 91–102. doi:10.1016/S0926-860X(02)00139-4.
  32. Yolcular, Sevim; Olgun, Özden (2008-11-01). "Ni/Al2O3 catalysts and their activity in dehydrogenation of methylcyclohexane for hydrogen production". Catalysis Today. Selected papers from the EUROPACAT VIII Hydrogen Society Session, Turku, Finland, 26–31 August 2007. 138 (3–4): 198–202. doi:10.1016/j.cattod.2008.07.020.
  33. Clot, Eric; Eisenstein, Odile; Crabtree, Robert H. "Computational structure–activity relationships in H2 storage: how placement of N atoms affects release temperatures in organic liquid storage materials". pubs.rsc.org. doi:10.1039/B705037B. Retrieved 2015-11-04.
  34. Eblagon, Katarzyna Morawa; Tam, Kin; Tsang, Shik Chi Edman. "Comparison of catalytic performance of supported ruthenium and rhodium for hydrogenation of 9-ethylcarbazole for hydrogen storage applications". Energy & Environmental Science. 5: 8621. doi:10.1039/C2EE22066K. Retrieved 2015-11-04.
  35. G. Laurenczy, C. Fellay, P. J. Dyson, Hydrogen production from formic acid. PCT Int. Appl. (2008), CODEN: PIXXD2 WO 2008047312 A1 20080424 AN 2008:502691
  36. Fellay, C; Dyson, PJ; Laurenczy, G (2008). "A Viable Hydrogen-Storage System Based On Selective Formic Acid Decomposition with a Ruthenium Catalyst". Angewandte Chemie International Edition in English. 47 (21): 3966–8. doi:10.1002/anie.200800320. PMID 18393267.
  37. F. Joó (2008). "Breakthroughs in Hydrogen Storage – Formic Acid as a Sustainable Storage Material for Hydrogen". ChemSusChem. 1 (10): 805–8. doi:10.1002/cssc.200800133. PMID 18781551.
  38. P. G. Jessop, in Handbook of Homogeneous Hydrogenation (Eds.: J. G. de Vries, C. J. Elsevier), Wiley-VCH, Weinheim, Germany, 2007, pp. 489–511.
  39. P. G. Jessop; F. Joó; C.-C. Tai (2004). "Recent advances in the homogeneous hydrogenation of carbon dioxide". Coordination Chemistry Reviews. 248 (21–24): 2425. doi:10.1016/j.ccr.2004.05.019.
  40. The ammonia economy. Memagazine.org (2003-07-10). Retrieved on 2012-01-08.
  41. Focus Denmark. Netpublikationer.dk (2006-06-13). Retrieved on 2012-01-08.
  42. Stracke, Marcelo P.; Ebeling, Günter; Cataluña, Renato; Dupont, Jairton (2007). "Hydrogen-Storage Materials Based on Imidazolium Ionic Liquids". Energy & Fuels. 21 (3): 1695–1698. doi:10.1021/ef060481t.
  43. Welch, G. C.; Juan, R. R. S.; Masuda, J. D.; Stephan, D. W. (2006). "Reversible, Metal-Free Hydrogen Activation". Science. 314 (5802): 1124–6. doi:10.1126/science.1134230. PMID 17110572.
  44. Elizabeth Wilson H2 Activation, Reversibly Metal-free compound readily breaks and makes hydrogen, Chemical & Engineering News November 20, 2006
  45. Mes stands for a mesityl substituent and C6F5 for a pentafluorophenyl group, see also tris(pentafluorophenyl)boron
  46. Graphene as suitable hydrogen storage substance. Physicsworld.com. Retrieved on 2012-01-08.
  47. Graphene to graphane. Rsc.org. January 2009. Retrieved on 2012-01-08.
  48. MOF-74 – A Potential Hydrogen-Storage Compound. Nist.gov. Retrieved on 2012-01-08.
  49. Researchers Demonstrate 7.5 wt% Hydrogen Storage in MOFs. Green Car Congress (2006-03-06). Retrieved on 2012-01-08.
  50. New MOF Material With hydrogen Uptake Of Up To 10 wt%. 22 February 2009
  51. Cella Energy
  52. Compendium of Hydrogen Energy.Volume 2:hydrogen Storage, Transportation and Infrastructure. A volume in Woodhead Publishing Series in Energy 2016,Chapter 8 – Other methods for the physical storage of hydrogen doi:10.1016/B978-1-78242-362-1.00008-0
  53. Stephen Lasher Analyses of Hydrogen Storage Materials and On-Board Systems. DOE Annual Merit Review June 7–11, 2010
  54. S&TR | Setting a World Driving Record with Hydrogen. Llnl.gov (2007-06-12). Retrieved on 2012-01-08.
  55. Compact (L)H2 Storage with Extended Dormancy in Cryogenic Pressure Vessels. Lawrence Livermore National Laboratory June 8, 2010
  56. Technical Sessions. FISITA 2010. Retrieved on 2012-01-08.
  57. Florusse, L. J.; Peters, CJ; Schoonman, J; Hester, KC; Koh, CA; Dec, SF; Marsh, KN; Sloan, ED (2004). "Stable Low-Pressure Hydrogen Clusters Stored in a Binary Clathrate Hydrate". Science. 306 (5695): 469–71. doi:10.1126/science.1102076. PMID 15486295.
  58. Zhevago, N.K.; Glebov, V.I. (2007). "Hydrogen storage in capillary arrays". Energy Conversion and Management. 48 (5): 1554–1559. doi:10.1016/j.enconman.2006.11.017.
  59. Zhevago, N.K.; Denisov, E.I.; Glebov, V.I. (2010). "Experimental investigation of hydrogen storage in capillary arrays". International Journal of Hydrogen Energy. 35: 169–175. doi:10.1016/j.ijhydene.2009.10.011.
  60. Dan Eliezer et al. A New Technology for Hydrogen Storage in Capillary Arrays. C.En & BAM
  61. Zhevago, N. K.; Chabak, A. F.; Denisov, E. I.; Glebov, V. I.; Korobtsev, S. V. (2013). "Storage of cryo-compressed hydrogen in flexible glass capillaries". International Journal of Hydrogen Energy. 38 (16): 6694–6703. doi:10.1016/j.ijhydene.2013.03.107.
  62. Glass microsphere diffusion. Ceer.alfred.edu (2001-05-15). Retrieved on 2012-01-08.
  63. G.G. Wicks; L.K. Heung; R.F. Schumacher. "SRNL's porous, hollow glass balls open new opportunities for hydrogen storage, drug delivery and national defense" (PDF). American Ceramic Society Bulletin. 87 (6): 23. Archived from the original (PDF) on November 15, 2008.
  64. R&D of large stationary hydrogen/CNG/HCNG storage vessels
  65. 1994 – ECN abstract. Hyweb.de. Retrieved on 2012-01-08.
  66. European Renewable Energy Network pp. 86, 188
  67. Hyunder
  68. Storing renewable energy: Is hydrogen a viable solution?
  69. Anscombe, Nadya (4 June 2012). "Energy storage: Could hydrogen be the answer?". Solar Novus Today. Retrieved 3 November 2012.
  70. Naturalhy
Wikimedia Commons has media related to Hydrogen storage.
This article is issued from Wikipedia - version of the 9/6/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.