Maëlle Cahuab,
Carlos A. Castilla-Martinezb,
Fabrice Salles
a,
Jérôme Long
ac and
Umit B. Demirci
*b
aICGM, Univ Montpellier, CNRS, ENSCM, Montpellier 34293, France
bInstitut Européen des Membranes, IEM – UMR 5635, Univ Montpellier, ENSCM, CNRS, Montpellier, France. E-mail: umit.demirci@umontpellier.fr
cInstitut Universitaire de France (IUF), 1 Rue Descartes, Paris Cedex 05 75231, France
First published on 31st July 2025
Ammonia borane (NH3BH3, referred to as AB) contains three protic and three hydridic hydrogen atoms, making it a promising candidate for solid-state hydrogen storage. However, in its pristine form, its practical application is limited by dehydrogenation temperatures exceeding 100 °C and the formation of byproducts with complex compositions. To overcome these limitations, we focused on destabilizing AB through confinement, made possible by using a Prussian blue analogue (PBA) as an oxygen-free host material. We selected the lacunar CoII[CoIII(CN)6]2/3□1/3 (referred to as CoCo) PBA due to its high thermal stability and the presence of coordinatively unsaturated Co2+ sites (CUS), which offer advantageous features for this purpose. Our results show that CoCo effectively confines AB, likely through a dual mechanism, with both chemisorption (where AB would coordinate to Co2+ CUS) and physisorption (where AB would be retained within the vacancies of the porous structure). Specifically, we found that approximately two-thirds of the AB would be chemisorbed, while one-third would be physisorbed. These findings highlight the crucial role of Co2+ CUS (of an oxygen-free host material) in AB confinement as well as in its isothermal dehydrogenation, likely involving homopolar B–H⋯H–B interactions.
In this regard, ammonia borane (AB) represents a promising molecular entity for chemical H2 storage owing to its high content in hydrogen (19.6 wt% H). However, the dehydrogenation of solid-state AB presents two key limitations that hinder its practical use for H2 storage: (i) moderately high desorption temperatures (100–200 °C) requiring additional energetic input for H2 release and (ii) a predominant decomposition that releases undesired volatile impurities (e.g. ammonia (NH3) and borazine (B3N3H6)) alongside H2.4,5 One strategy to address these issues involves confining AB within porous materials to facilitate the dehydrogenation process.6 In particular, previous studies have focused on confining AB within porous molecule-based materials such as metal–organic frameworks (MOF; see e.g. ref. 7–16). While a reduction in the dehydrogenation temperature of confined AB is observed, the underlying mechanisms remain unclear. These effects may arise from enhanced surface tension generated by the MOF microporosity or from a catalytic effect mediated by coordinatively unsaturated sites (CUS) within the framework. Additionally, coordinated solvents on the CUS might play a significant role.17 Gaining further fundamental understanding of the mechanism involved in AB confinement requires exploration of alternative materials. It is worth mentioning that O-containing linkers in MOFs such as Zn-MOF-74 (ref. 18) and Fe-MIL-53 (ref. 19) destabilize AB by reaction of O of the linker and BH3 of AB, resulting in B–O bonds.
To address the challenge of limiting the decomposition of confined AB, it is critical to develop porous frameworks that are free of hydrogen and oxygen species, which include cyano-bridged coordination networks. Prussian Blue Analogues (PBA), which are typical representatives of cyano-bridged coordination networks, offer intriguing possibilities that remain largely unexplored for H2 storage applications.20 PBA exhibits the general formula AaM[M′(CN)6]b□c·xH2O (where A is an alkali ion, M and M′ are transition metal ions, and □ represents the cyanometallate vacancies that ensure the electroneutrality). In these structures, octahedral [M′(CN)6]x− complexes are linked through cyano bridges to octahedral Mn+ ions, creating a 3D cubic porous structure (pore sizes of 4 Å and 7.5 Å). The presence of cyanometallate vacancies indicates that the Mn+ cation completes its coordination sphere with water molecules, which can be subsequently removed to create CUS, able to interact with various molecular species. Moreover, the nature of the metal ions could be easily modulated without affecting the resulting cubic porous structures, allowing for the accurate determination of the parameters affecting the sorption features. Hence, these porous networks have been investigated for sorbents against water,21 ammonia,22 carbon dioxide,23 sulfur oxide and hydrogen sulfide,24 methane,25 propane/xylene26 and hydrocarbons,27 as well as H2.28,29 For the latter, despite pioneering studies, the enhancement of the H2 storage features family has been rather limited due to their relatively moderate storage capacity (2–3 wt% H2) at −196 °C by physisorption measurements. However, upon thermal activation and water removal, PBAs exhibit frameworks free of O and H atoms, an essential feature for stable and efficient AB adsorption and confinement.
To our knowledge, the use of PBA for confining AB and facilitating its controlled H2 release remains unexplored. In this study, we investigate the potential of PBA as materials for AB confinement and study their role in destabilization within the coordination network. Specifically, we focus on the lacunary CoII[CoIII(CN)6]2/3□1/3 (denoted CoCo) as a model system due to its high thermal stability and the presence of Co2+ CUS that can react with the nitrogen of AB via a coordination bond. Our findings demonstrate that CoCo exhibits remarkable capacity for AB confinement, and reveals a dual adsorption mechanism involving both chemisorption (AB coordinated to Co2+ CUS) and physisorption (AB confined within the PBA's porosity). Notably, we quantified the proportions of each adsorbed AB species, highlighting the crucial role of Co2+ CUS in the confinement process. These insights not only establish a new approach to AB confinement, but also provide a foundation for optimizing hydrogen storage systems by modulating the metal ion nature or employing post-functionalization strategies.
An activation phase followed,27 during which CoCo·H2O was transferred to a flask equipped with a cap fitted with a needle. The setup was inserted into a Schlenk line and placed under vacuum while heating at 160 °C for 12 h. During this process, the color of the powder changed. The activated PBA, denoted as CoCo hereafter, exhibits a royal blue color (Fig. 2). PBA was then stored in an argon-filled glovebox (O2 and H2O: <0.1 ppm) to protect it from exposure to water and prevent its rehydration.
As mentioned above and as detailed in the next main section, we then aimed at distinguishing the fraction of AB that is chemisorbed onto Co2+ CUS and the fraction that is physisorbed (i.e. AB confined) within the PBA's porosity. Typically, in a flask and under argon, CoCo·AB (33 mg) was dispersed in diethyl ether solvent (6 mL) and the suspension was placed under magnetic stirring. After 24 h, the stirring was stopped, and the supernatant, having extracted the AB molecules weakly attached to the PBA network, was recovered to be analyzed. The as-obtained sample is denoted as CoCo·AB*. It was finally subjected to a two-day vacuum treatment to remove the solvent.
Liquid-state 11B nuclear magnetic resonance spectroscopy (NMR; B0(1H) 400 MHz; B0(11B) 128.4 MHz) was conducted on AB solutions (see hereafter) using DMSO-d6 to lock the signal (Sigma-Aldrich, >99%) in a capillary tube. Anhydrous dichloromethane or anhydrous diethyl ether was employed as the solvent, depending on the sample being analyzed. It is worth mentioning that the paramagnetic nature of CoII has not allowed for recording exploitable 11B MAS NMR spectra of the aforementioned solids.
Structural analyses were conducted using powder X-ray diffraction (PXRD). X-ray patterns were acquired with the PANalytical Empyrean M diffractometer in Bragg configuration, covering a range from 10° to 60° in 2θ. The step size was set to 0.026° in 2θ, and each step had a duration of 2700 s. A Co source with a Kα1 wavelength of 1.788965 Å was utilized.
The textural properties of CoCo were determined by N2 sorption using a Micromeritics TriFlex apparatus. The specific surface area and the total pore volume were calculated by the Brunauer–Emmett–Teller BET and the Barrett, Joyner and Halenda BJH methods. As both bulk AB and confined AB dehydrogenate upon heating at isothermal conditions, for example at temperatures as low as 70 and 40 °C, respectively, measurements of the textural properties for CoCo·AB were not feasible. Indeed, preparing the sample for analysis involves degassing under heating, and this leads to the dehydrogenation of AB.
Energy-dispersive X-ray spectroscopy (EDS) was carried out. The measurements were performed on a scanning electron microscope Quanta 200 Hitachi S-2600N equipped with an Oxford Instruments X-Max 50 mm2 detector. Element quantification was performed using the INCA software with an acquisition time of 30 s, and all measurements conducted under vacuum conditions. Elemental analyses were conducted using an Elementar Vario MICRO Cube analyzer, with gas chromatography columns developed by Elementar. Additionally, inductively coupled plasma-mass spectroscopy (ICP-MS) analyses were performed using an Agilent 7900 machine. For boron analysis, the machine was operated in “no gas” mode with an analysis time of 0.1 s, and five measurements were taken.
Thermogravimetric and differential thermal analyses (TGA and DTA) were conducted using the thermal analyzer SDT Q600 TA Instruments. The heating process ranged from 25 °C with a heating rate of 5 °C per minute and a N2 flow of 100 mL min−1. The released gases were analyzed by TG-coupled mass spectrometry (MS; QMS 403 Aëolos Quadro). A hermetically sealed aluminum crucible was used to prevent the samples from exposure to ambient air and moisture, and the crucible lid was pierced just before the analysis to allow the escape of gases.
After the isothermal treatment, the atmosphere inside the reactor was analyzed by gas chromatography (GC) using a PerkinElmer Clarus 400 apparatus equipped with a ShinCarbon ST column. A 100 μL sample was injected into the GC to quantify the H2 content. This analysis was performed four times to confirm reproducibility. The solid obtained after the isothermal experiment was recovered and analyzed by FTIR spectroscopy.
![]() | ||
Fig. 3 FTIR spectra of CoCo·H2O, CoCo and CoCo·AB. For comparative purpose, the spectra for free AB is also given. |
The PXRD pattern of CoCo·H2O (Fig. 4) matches with the face-centered cubic (fcc) structure with a space group Fmm (225), consistent with the well-known fcc crystallographic structure of PBA, where the Co2+ and Co3+ ions are connected through the cyano-bridge forming a 3D cubic structure. Using the Bragg's law and based on the (2 0 0) reflection, the cell parameter a was found to be 10.27 Å.
![]() | ||
Fig. 4 PXRD for CoCo·H2O, CoCo and CoCo·AB. The right part represents the magnification of the (2 0 0) peak. |
To favor AB adsorption on the Co2+ sites of the PBA host material, removal of the water molecules was achieved by heating CoCo·H2O at 160 °C for 12 h under vacuum, yielding the activated lacunar PBA CoII[CoIII(CN)6]2/3□1/3, also denoted as CoCo. This process intends to generate Co2+ CUS that are highly reactive towards the nitrogen of AB. The activation was visually confirmed by a color change from pink to royal blue (Fig. 2). The blue color likely arises from a minor fraction of Co2+ ions adopting a four-coordinated environment. Due to their much higher molar extinction coefficient compared to octahedral Co2+, this minor fraction can dominate the optical response. Moreover, the FTIR spectrum and the TGA curve of CoCo showed significant attenuation of the O–H bands (Fig. 3) and the absence of an abrupt weight up to 300 °C (Fig. S2 in comparison to those of CoCo·H2O). The integrity of PBA upon activation was confirmed by the persistence of the CN band at 2173 cm−1 in the FTIR spectrum (Fig. 3) and the retention of the fcc structure (cell parameter a = 10.09 Å based on the (2 0 0) reflection), as evidenced by the PXRD pattern (Fig. 4). Due to limitations in our experimental setup, it is very challenging to completely isolate the sample from air during transfer from the glovebox to the analysis instruments. This minimal exposure likely explains the presence of weak O–H bands in the FTIR spectrum and the slight deviation (weight loss of about 1 wt% at about 300 °C) observed in the TGA curve.
The textural properties of CoCo were determined by N2 sorption (Fig. S3). According to the IUPAC classification, the adsorption isotherm is of type IV, typical of a mesoporous material. The specific surface area of CoCo was found to be as high as 903.6 m2 g−1. Its total pore volume is 0.80 cm3 g−1; assuming that the internal volume of CoCo can ideally be fully filled by AB (780 mg cm−3). This will require a loading of approximately 620 mg of AB (per gram of CoCo).
CoCo·AB was obtained through solubilization of AB in anhydrous dichloromethane and reaction with PBA using a 1:
1 molar ratio. To assess the stability of AB during the impregnation process, the supernatant was sampled and analyzed by 11B NMR. The spectrum (Fig. 4) showed a single signal consisting of a 1
:
3
:
3
:
1 quartet with a chemical shift δ of −22 ppm and a coupling constant 1JB–H of 97 Hz. This is consistent with a BH3 group and can be viewed as a signature of AB.36 All of these indicated that AB in dichloromethane is stable during the impregnation of PBA. Otherwise, oligomerization of AB would have resulted in a triplet signal at around −13 ppm due to the presence of BH2 environments.37
The successful impregnation of AB in PBA, forming CoCo·AB, was evident from a color shift from royal blue to dark purple (Fig. 2). This observation would suggest a change in the coordination sphere of the Co2+ ion. Yet, analysis of the color change is beyond the scope of this work due to the broad nature of the d–d absorption bands involved. Moreover, the change in the coordination sphere might only affect a subset of the Co2+ ions.
The FTIR spectrum of CoCo·AB (Fig. 3) confirms the structural stability of the PBA network after impregnation with AB, as indicated by the persistence of the CN band at 2172 cm−1. Additionally, the spectrum exhibits the main vibration bands characteristic of pure AB, supporting the successful impregnation process. The bands are of weak intensity, which is generally observed for AB confined in the porosity of a host material.7 Interestingly, a broadening of the N–H deformation bands (1700–1300 cm−1) is observed. This broadening might be attributed to a change in the local environment experienced by the NH3 group of AB38 after impregnation within the CoCo network.
The PXRD pattern of CoCo·AB (Fig. 4) confirmed the integrity of the fcc structure of the host material, indicating that the network remained intact upon AB impregnation. The absence of diffraction peaks related to crystalline AB suggests that AB is located within the porosity of the CoCo network as AB in this state does not diffract. This is in line with previous observations on confined AB. Using Bragg's law, the estimated cell parameter a of PBA was determined to be 10.20 Å. This value is slightly lower than that of pristine CoCo·H2O (a = 10.27 Å), suggesting that the volume occupied by AB molecules within the porosity is lower than the volume occupied by the water molecules in the as-synthesized CoCo·H2O.
The loading of AB in PBA was further confirmed by quantifying the boron content in CoCo·AB using ICP-MS. The measured weight percentage of boron was found to be 4.5 wt%, which translates to a weight percentage of 12.85 wt% for AB. This also corresponds to an AB:
CoCo molar ratio of approximately 0.97
:
1.
The remaining boron content in CoCo·AB* was determined by ICP-MS. It corresponded to 3.1 wt% B, which indicates a weight percentage of AB of 8.8 wt%. In other words, approximately two thirds of the impregnated AB remained within the CoCo network, while one-third was removed. This suggests that about two-thirds of the AB molecules would have coordinated to the Co2+ CUS, while the remaining one-third would have interacted weakly with the framework. For CoCo·AB*, the (experimental) molar ratio between the coordinated AB and CoCo is approximately 0.63:
1, suggesting the formula CoII[CoIII(CN)6]2/3□1/3·0.63NH3BH3. This suggests that almost all of the Co2+ CUS would be coordinated to one AB molecule.
![]() | ||
Fig. 5 TGA curve for CoCo·AB* under argon, and MS analysis (m/z = 2 for H2, m/z = 17 for NH3 and m/z = 29, 31 and 45) for (C2H5)2O. |
Under isothermal conditions (Fig. 6a), CoCo·AB* was dehydrogenated in a single step, with the extent of the dehydrogenation increasing as the temperature rose. The amount of gas released at 180 min provides a point of comparison: whereas pristine AB released 0.83 equiv. H2 at 85 °C, CoCo·AB* produced 0.44, 0.67, 0.84 and 1.02 equiv. H2 at 60, 70, 80 and 90 °C, respectively. The duration of the dehydrogenation process was significantly shorter for CoCo·AB*, taking approximately 40, 30, 25, and 20 minutes at 60, 70, 80, and 90 °C, respectively. This is a notable improvement compared to pristine AB, which required around 150–200 minutes at 80–90 °C according to references,44–46 and 180 minutes at 85 °C under our experimental conditions (curve not reported). All of these demonstrate the successful destabilization of AB coordinated to the Co2+ CUS. Indeed, the dehydrogenation process is significantly accelerated, completing in just 20–25 minutes within the same temperature range. This translates to a 6- to 9-fold increase in the dehydrogenation kinetics (Fig. 6b). The dehydrogenation curves (Fig. 6a) were exploited to determine the apparent activation energy using the Arrhenius equation (Fig. S7), yielding a value of 60 kJ mol−1. This value is much lower than the 183 kJ mol−1 reported for pristine AB,47 and is comparable to the 71 kJ mol−1 reported for e.g. lithium amidoborane LiNH2BH3, a metal derivative of AB.48 These findings further confirm the destabilization of confined AB.
![]() | ||
Fig. 6 FTIR spectra of the solids recovered after dehydrogenation of CoCo·AB* under isothermal conditions. The bands have been assigned in the colored rectangles. |
The CoCo·AB* samples recovered after the isothermal dehydrogenation experiments exhibited a purple color (Fig. 2) and were all analyzed by FTIR spectroscopy (Fig. 7). The presence of residual diethyl ether was confirmed by the C–H stretching bands at around 2800 cm−1. The CoCo network remained intact in the dehydrogenation temperature range of 60–90 °C, as indicated by the unchanged CN stretching vibrational mode. The B–H stretching mode of AB (2200–2600 cm−1) decreases in intensity as the temperature increases, disappearing at 90 °C. In contrast, the N–H stretching mode (3100–3500 cm−1) also decreased in intensity but remained detectable, even at 90 °C. The disappearance of the B–H stretching mode at 90 °C, alongside the persistence of the N–H stretching mode, suggests that homopolar B–H⋯H–B dehydrogenation49 plays a role in the dehydrogenation of CoCo·AB*.
An additional experiment was conducted by heating CoCo·AB* at 150 °C overnight. The recovered solid was analyzed by FTIR spectroscopy (Fig. 8). The dehydrogenated sample still exhibited the N–H stretching vibrational mode, while the B–H stretching mode had disappeared. There were, however, notable changes in the N–H bending mode (1300–1700 cm−1): the band at approximately 1400 cm−1 became sharper and more defined, while the bands between 1000 and 1300 cm−1 broadened and merged. This last region is associated with the stretching vibrational mode of Co−NHx.50
![]() | ||
Fig. 8 FTIR spectra of pristine CoCo·AB*, CoCo·NH3 and the respective solids recovered after isothermal treatment at 150 °C. |
To further explore this, CoCo (i.e. free of any guest molecules like both H2O and AB) was saturated with NH3 under a stream of N2 and NH3, resulting in the sample denoted CoCo·NH3 (Fig. 2), as confirmed by FTIR spectroscopy (Fig. 8). CoCo·NH3 was then treated at 150 °C under isothermal conditions, and its FTIR spectra were compared with that of CoCo·AB* dehydrogenated at the same temperature (Fig. 8). Both spectra exhibited the band associated with Co–NHx at about 1400 cm−1 in the N–H bending region, as well as a small band at around 3650 cm−1 in the N–H stretching region (observable in neither CoCo·NH3 nor CoCo·AB* before heating). It is reasonable to conclude that under heating, the likely interaction of NHx from AB with the Co2+ CUS would be more stable than the BHy moiety, with the latter undergoing more dehydrogenation than the former. A last observation from the FTIR spectrum of CoCo·AB* dehydrogenated at 150 °C is that the B–N stretching mode is present, suggesting the formation of the [NHxBHy]n species with x > y and n ≥ 1.
When exposed for a duration of 4 hours to flowing air (200 mL min−1) with controlled humidity (42%) and at 25 °C, CoCo·AB* evolved. Its color changed from deep purple to dark pink, distinct from the bright pink color of CoCo·H2O (Fig. 2). The FTIR spectrum (Fig. 9) resembled a combination of the spectra of ‘fresh’ CoCo·AB* and CoCo·H2O. This suggests that partial substitution of AB coordinated to the Co2+ CUS by H2O occurred. CoCo·AB* is thus not stable in the presence of moisture, indicating that H2O is a stronger Lewis base than AB. The N–B bond in AB seems to weaken the Lewis basicity of NH3 in AB, which contrasts with the well-known stronger basicity of free NH3 compared to H2O in coordination chemistry. Furthermore, the presence of B–O bonds in the 1000–1200 cm−1 region of the spectrum cannot be ruled out, likely due to the rapid hydration and/or hydrolysis of the BH3 moiety of the AB molecule.51 The persistence of N–H vibrational modes suggests that the substitution by water was not complete within the 4 hours exposure period.
We also provide further insights into the hydrogen release mechanism, with the likely occurrence of homopolar B–H⋯H–B dehydrogenation leading to the [NHxBHy]n species with x > y and n ≥ 1. This offers promising perspectives to modulate the nature of the transition metal ions or using a post-functionalization rationale to optimize the AB confinement and control its dehydrogenation.
Fig. S1–S8. See DOI: https://doi.org/10.1039/d5se00758e.
This journal is © The Royal Society of Chemistry 2025 |