Abstract
Hydrogen bonds profoundly influence the fundamental chemical, physical and biological properties of molecules and materials. Owing to their relatively weaker interactions compared to other chemical bonds, hydrogen bonds alone are generally insufficient to induce substantial changes in electrical properties, thus imposing severe constraints on their applications in related devices. Here we report a metal–insulator transition controlled by hydrogen bonds for an organic–inorganic (1,3-diaminopropane)0.5SnSe2 superlattice that exhibits a colossal on–off ratio of 107 in electrical resistivity. The key to inducing the transition is a change in the amino group’s hydrogen-bonding structure from dynamic to static. In the dynamic state, thermally activated free rotation continuously breaks and forms transient hydrogen bonds with adjacent Se anions. In the static state, the amino group forms three fixed-angle positions, each separated by 120°. Our findings contribute to the understanding of electrical phenomena in organic–inorganic hybrid materials and may be used for the design of future molecule-based electronic materials.
This is a preview of subscription content, access via your institution
Access options
Access Nature and 54 other Nature Portfolio journals
Get Nature+, our best-value online-access subscription
$29.99 / 30 days
cancel any time
Subscribe to this journal
Receive 12 print issues and online access
$259.00 per year
only $21.58 per issue
Buy this article
- Purchase on SpringerLink
- Instant access to full article PDF
Prices may be subject to local taxes which are calculated during checkout
Similar content being viewed by others
Data availability
The data that support the findings of this study are available within the Supplementary Information. Source data are provided with this paper.
References
Pimentel, G. C. & McClellan, A. L. Hydrogen bonding. Annu. Rev. Phys. Chem. 22, 347–385 (1971).
Kollman, P. A. & Allen, L. C. Theory of the hydrogen bond. Chem. Rev. 72, 283–303 (1972).
Steiner, T. The hydrogen bond in the solid state. Angew. Chem. Int. Ed. 41, 48–76 (2002).
Buemi, G. in Hydrogen Bonding—New Insights (ed. Grabowski, S. J.) 51–107 (Springer, 2006).
Aakeröy, C. B. & Seddon, K. R. The hydrogen bond and crystal engineering. Chem. Soc. Rev. 22, 397–407 (1993).
Jeffrey, G. A. & Saenger, W. Hydrogen Bonding in Biological Structures (Springer, 1991).
Doster, W. The dynamical transition of proteins, concepts and misconceptions. Eur. Biophys. J. 37, 591–602 (2008).
Dahl, P. J. et al. A 300-fold conductivity increase in microbial cytochrome nanowires due to temperature-induced restructuring of hydrogen bonding networks. Sci. Adv. 8, eabm7193 (2022).
Portfield, W. W. (ed.) in Inorganic Chemistry 2nd edn, 885–893 (Academic, 1993).
Mott, N. F. Metal–insulator transition. Rev. Mod. Phys. 40, 677–683 (1968).
Yang, Z., Ko, C. & Ramanathan, S. Oxide electronics utilizing ultrafast metal-insulator transitions. Annu. Rev. Mater. Res. 41, 337–367 (2011).
Wang, Z. et al. Resistive switching materials for information processing. Nat. Rev. Mater. 5, 173–195 (2020).
Asamitsu, A., Tomioka, Y., Kuwahara, H. & Tokura, Y. Current switching of resistive states in magnetoresistive manganites. Nature 388, 50–52 (1997).
Qazilbash, M. M. et al. Mott transition in VO2 revealed by infrared spectroscopy and nano-imaging. Science 318, 1750–1753 (2007).
Liang, Y. G. et al. Tuning the hysteresis of a metal-insulator transition via lattice compatibility. Nat. Commun. 11, 3539 (2020).
Tian, N. et al. Dimensionality-driven metal to Mott insulator transition in two-dimensional 1T-TaSe2. Natl Sci. Rev. 11, nwad144 (2024).
Zhang, W. et al. Role of vacancies in metal–insulator transitions of crystalline phase-change materials. Nat. Mater. 11, 952–956 (2012).
Chu, C. W., Harper, J. M. E., Geballe, T. H. & Greene, R. L. Pressure dependence of the metal-insulator transition in tetrathiofulvalinium tetracyanoquinodimethane (TTF-TCNQ). Phys. Rev. Lett. 31, 1491–1494 (1973).
Lee, S.-H., Goh, J. S. & Cho, D. Origin of the insulating phase and first-order metal-insulator transition in 1T-TaS2. Phys. Rev. Lett. 122, 106404 (2019).
Lu, N. et al. Electric-field control of tri-state phase transformation with a selective dual-ion switch. Nature 546, 124–128 (2017).
Li, L. et al. Manipulating the insulator–metal transition through tip-induced hydrogenation. Nat. Mater. 21, 1246–1251 (2022).
Yi, W. et al. Biological plausibility and stochasticity in scalable VO2 active memristor neurons. Nat. Commun. 9, 4661 (2018).
Ueda, A. et al. Hydrogen-bond-dynamics-based switching of conductivity and magnetism: a phase transition caused by deuterium and electron transfer in a hydrogen-bonded purely organic conductor crystal. J. Am. Chem. Soc. 136, 12184–12192 (2014).
Isono, T. et al. Hydrogen bond-promoted metallic state in a purely organic single-component conductor under pressure. Nat. Commun. 4, 1344 (2013).
Zhou, J. et al. Layered intercalation materials. Adv. Mater. 33, 2004557 (2021).
Passarelli, J. V. et al. Tunable exciton binding energy in 2D hybrid layered perovskites through donor–acceptor interactions within the organic layer. Nat. Chem. 12, 672–682 (2020).
Jin, S. et al. High-Tc superconducting phases in organic molecular intercalated iron selenides: synthesis and crystal structures. Chem. Commun. 53, 9729–9732 (2017).
Zhang, H. et al. Enhancement of superconductivity in organic-inorganic hybrid topological materials. Sci. Bull. 65, 188–193 (2020).
Coronado, E. et al. Coexistence of superconductivity and magnetism by chemical design. Nat. Chem. 2, 1031–1036 (2010).
Wan, C. et al. Flexible n-type thermoelectric materials by organic intercalation of layered transition metal dichalcogenide TiS2. Nat. Mater. 14, 622–627 (2015).
Wang, N. et al. Transition from ferromagnetic semiconductor to ferromagnetic metal with enhanced Curie temperature in Cr2Ge2Te6 via organic ion intercalation. J. Am. Chem. Soc. 141, 17166–17173 (2019).
Ying, T. P. et al. Observation of superconductivity at 30∼46 K in AxFe2Se2(A = Li, Na, Ba, Sr, Ca, Yb and Eu). Sci. Rep. 2, 426 (2012).
Ying, T. et al. Superconducting phases in potassium-intercalated iron selenides. J. Am. Chem. Soc. 135, 2951–2954 (2013).
Sun, R. et al. Intercalating anions between terminated anion layers: unusual ionic S–Se bonds and hole-doping induced superconductivity in S0.24(NH3)0.26Fe2Se2. J. Am. Chem. Soc. 141, 13849–13857 (2019).
Fan, X. et al. Nematicity and superconductivity in orthorhombic superconductor Na0.35(C3N2H10)0.426Fe2Se2. Phys. Rev. Mater. 2, 114802 (2018).
Kobayashi, M. & Tanaka, H. The reversibility and first-order nature of liquid–liquid transition in a molecular liquid. Nat. Commun. 7, 13438 (2016).
Luo, J.-B., Wei, J.-H., Zhang, Z.-Z., He, Z.-L. & Kuang, D.-B. A melt-quenched luminescent glass of an organic–inorganic manganese halide as a large-area scintillator for radiation detection. Angew. Chem. Int. Ed. 62, e202216504 (2023).
Benoit, M., Marx, D. & Parrinello, M. Tunnelling and zero-point motion in high-pressure ice. Nature 392, 258–261 (1998).
Goncharov, A. F. et al. Dynamic ionization of water under extreme conditions. Phys. Rev. Lett. 94, 125508 (2005).
Ye, Q.-J., Zhuang, L. & Li, X.-Z. Dynamic nature of high-pressure ice VII. Phys. Rev. Lett. 126, 185501 (2021).
Hedges, L. O., Jack, R. L., Garrahan, J. P. & Chandler, D. Dynamic order-disorder in atomistic models of structural glass formers. Science 323, 1309–1313 (2009).
Hernandez, J.-A. & Caracas, R. Superionic–superionic phase transitions in body-centered cubic H2O-ice. Phys. Rev. Lett. 117, 135503 (2016).
Reinhardt, A. et al. Thermodynamics of high-pressure ice phases explored with atomistic simulations. Nat. Commun. 13, 4707 (2022).
Tayran, C. & Çakmak, M. Charge density wave in a SnSe2 layer on and the effect of surface hydrogenation. Phys. Chem. Chem. Phys. 24, 6820–6827 (2022).
Smith, A. J., Meek, P. E. & Liang, W. Y. Raman scattering studies of SnS2 and SnSe2. J. Phys. C Solid State Phys. 10, 1321 (1977).
Kohn, W. & Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 140, A1133–A1138 (1965).
Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996).
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).
Porezag, D. & Pederson, M. R. Infrared intensities and Raman-scattering activities within density-functional theory. Phys. Rev. B 54, 7830–7836 (1996).
Grimme, S., Bannwarth, C. & Shushkov, P. A robust and accurate tight-binding quantum chemical method for structures, vibrational frequencies and noncovalent interactions of large molecular systems parametrized for all spd-block elements (Z = 1–86). J. Chem. Theory Comput. 13, 1989–2009 (2017).
Kühne, T. D. et al. CP2K: an electronic structure and molecular dynamics software package - Quickstep: efficient and accurate electronic structure calculations. J. Chem. Phys. 152, 194103 (2020).
Bussi, G., Donadio, D. & Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 126, 014101 (2007).
Acknowledgements
We thank Y. P. Song, Y. T. Song, S. F. Jin, J. Y. Zhou and R. J. Sun for discussions. This work was financially supported by the National Key Research and Development Program of China (grants 2021YFA1401800 and 2022YFA1203200), Beijing Natural Science Foundation (grant Z200005) and the National Natural Science Foundation of China (grants 52272267, 52202342, 51922105, 12022412 and 22273014). This work was also supported by the Synergetic Extreme Condition User Facility (SECUF), including the THz and Infrared Experimental Station, Infrared Unit and Sample Preselection and Characterization Station. A portion of numerical computations were carried out at the Hefei Advanced Computing Center.
Author information
Authors and Affiliations
Contributions
T.Y., Y.G., J.G. and X.C. conceptualized the idea and provided project supervision. Z.X., T.Y., B.S. and X.C. synthesized the samples. Z.X., T.Y., M.H., C.C., J.Y. and Z.C. conducted the physical property characterizations and data processing. R.L., Y.G., C.Z. and T.Y. performed the theoretical calculations. Z.H. and L.D. measured the Raman spectra. T.Y. and Y.G. analysed the data and wrote the paper, with input from all authors.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Chemistry thanks Xia Wang, Nikhil Malvankar, Peter Dahl and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Extended data
Extended Data Fig. 1 SnSe2 and (1,3-Diaminopropane)0.5SnSe2 single crystals.
Optical images of the single crystals (a) before and (b) after intercalations. c. X-ray diffraction patterns of pristine SnSe2 and intercalated (1,3-DAP)0.5SnSe2 with preferred (00 l) diffractions.
Extended Data Fig. 2 Temperature-dependent single crystalline X-ray diffraction.
a, b. Temperature-dependent lattice parameters a, b, c, V of (1,3-DAP)0.5SnSe2 in the temperature range 100 K- 400 K. The space group of (1,3-DAP)0.5SnSe2 is automatically determined to be C2/m (No.12) in the whole temperature range without the observation of structural transition around 150 K. The refinement error bars have been incorporated. c, d. Single crystal XRD diffraction spots showing hk0 reciprocal lattice planes at 300 K and 100 K, respectively. The magnified view of selected diffraction points in both linear and logarithmic scales show that no superstructure diffraction peaks can be observed at 100 K.
Extended Data Fig. 3 Structure of the lowest-energy configuration of (1,3-DAP)0.5SnSe2.
a, b. top view, and c. side view of (1,3-DAP)0.5SnSe2, showing that alternative two rows of 1,3-DAP molecules tilted at an angle of 120°. Blue and red dashed double lines in c. represent two sets of hydrogen bond interactions between 1,3-DAP and Se atoms, and between the intermolecular amino groups (-NH2), respectively.
Extended Data Fig. 4 Temperature dependence of resistivity.
a. Temperature dependence of resistivity for different (1,3-DAP)0.5SnSe2 samples, noting that the resistivity exhibits violent fluctuations above room temperature (indicated by light purple area) due to the motion of interlayer organic molecules during melting-freezing transition. b. The temperature dependence of dρ/dT of (1,3-DAP)0.5SnSe2 sample #1-#6 and #6*, which demonstrates the same TD of 160 K.
Extended Data Fig. 5 Working mechanism of the dissolved water.
a. Chemical equations of the protonated ammonia molecule and 1,3-DAP molecule. b. The pH values of Ultra-dry 1,3-DAP in the glove box, showing a PH value of 6.8. c. The pH values of water-contained 1,3-DAP in open air, showing a PH value of 13.7. d. Electrical resistance of 1,3-DAP precursor with the variation of dissolved water content. e. Co-intercalation of Rb metal and 1,3-DAP into SnSe2, leading to substantial electron doping that overshadows charge transfer from the dynamic-static transition of hydrogen bonds. Note that the commercially purchased 1,3-DAP contains trace amounts of dissolved water where the ionized hydroxide ions serve as electron donors, mirroring the electron-doping role of Rb metal co-intercalated in Rbx(1,3-DAP)ySnSe2. f. A numerical simulation employing a simple parallel circuit model based on data from panels h (1000 layers) and e (1 layer), visually demonstrating the observed decreasing tendency (with decreasing temperature) in the high-R region. g. Logarithmic plot of the resistivity of (1,3-DAP)0.5SnSe2 with trace amounts of water. h. Temperature-dependent resistivity of the (1,3-DAP)0.5SnSe2 single crystal, prepared using molecular sieve-treated ultra-dry 1,3-DAP solvent, revealing a notable absence of metallic behavior at low temperature for the pure sample.
Extended Data Fig. 6 Physical properties of (EDA)xSnSe2 superlattice.
a. Optical images of SnSe2 and (EDA)xSnSe2 single crystals. b. Energy dispersive X-ray (EDX) mapping of the acquired (EDA)xSnSe2 single crystal. c. X-ray diffraction patterns of (EDA)xSnSe2. d. Temperature-dependent resistivity of two batches of (EDA)xSnSe2s single crystals with different dissolved water content in the EDA solvent.
Extended Data Fig. 7 PXRD of Sn(Se1-xSx)2 and (1,3-DAP)0.5Sn(Se1-xSx)2 crystals, and Stability of the MIT.
The X-ray diffraction patterns of a. SnSe2-xSx (x = 0,0.071,0.101,0.296) and c. (1,3-DAP)0.5SnSe2-xSx (x = 0,0.071,0.101,0.296), showing a series of (00 l) diffractions. Enlarge view of (001) peaks are displayed in b and d. e. Different runs of resistance measurements of (1,3-DAP)0.5SnSe2 (sample #C) by repeated heating and cooling with the highest temperature below 300 K. f. Different runs of resistance measurements of a S-doped sample (sample #D). g. Resistance of the air-exposed sample #C for one week.
Extended Data Fig. 8 Dynamics of amino groups.
a, b. Calculated θ values of representative –NH2 groups in the simulation system, based on 20-ps NPT simulations of (1,3-DAP)0.5SnSe2 at temperatures of 300 and 100 K respectively. At 300 K, θ curves show pronounced fluctuations (steps and hops), exhibiting that -NH2 groups keep self-rotating. In contrast, at 100 K, θ values show negligible variations, implying that the -NH2 groups no long self-rotate. c. Temperature evolution during the MD cooling process, ranging from 400 to 100 K, showing a linear correlation between temperature and simulation time. d. Angular variations (θ) of the -NH2 groups throughout the cooling simulation. The region highlighted by the red line frame exhibits frequent and large fluctuations of θ, indicating frequent -NH2 self-rotation at high temperature, while the region highlighted by the blue line frame shows θ with negligible fluctuations, indicating -NH2 stopping self-rotation as temperature decreases to low temperature.
Extended Data Fig. 9 Second-order phase transition of (1,3-DAP)0.5SnSe2.
a, b. Experimentally measured resistivity during heating (red line) / cooling (blue line) processes of the ultra-try and water-contained (1,3-DAP)0.5SnSe2 single crystals, respectively. c. Interposition hopping rate ν obtained from MD simulations during heating (red line) / cooling (blue line) processes.
Extended Data Fig. 10 Electronic structures before and after 1,3-DAP insertion.
a. Projected density of states (PDOS) of SnSe2, showing Eg = 0.71 eV. b. PDOS of (1,3-DAP)0.5SnSe2, showing Eg extremely close to 0 eV. c. Band gap at different rotation angles. Eg of (1,3-DAP)0.5SnSe2 configurations with one representative -NH2 at varied self-rotation angles, showing Eg varying within 0.06 eV, extremely close to 0 eV. d. Effective electronic delocalization. Number of partial charge on the first band ‘below’ EF on representative molecules in the same layer when one -NH2 self-rotates at different θ values. It is noticeable that charge on one -NH2 located at various positions varies with θ, equivalent to charge delocalization.
Supplementary information
Supplementary Information
Supplementary Tables 1 and 2
Supplementary Video 1
-NH2 rotation in (DAP)0.5SnSe2 at 300 K
Supplementary Video 2
-NH2 vibration in (DAP)0.5SnSe2 at 100 K
Supplementary Video 3
-NH2 behaviour in (DAP)0.5SnSe2 during cooling
Supplementary Video 4
Charge redistribution of HOMO as one -NH2 group rotates
Supplementary Video 5
Charge redistribution of HOMO as one -NH2 group rotates (a larger cell)
Supplementary Video 6
Raman vibrational mode at 251 cm−1 calculated by DFT
Source data
Source Data Fig. 1
Statistical source data
Source Data Fig. 2
Statistical source data
Source Data Fig. 3
Statistical source data. Input files and scripts for DFT calculations using VASP or CP2K
Source Data Fig. 4
Statistical source data. Input files, some output files and scripts for DFT calculations using VASP or CP2K crystal structure of vesta file
Source Data Extended Data Fig. 1
Unprocessed raw test data
Source Data Extended Data Fig. 2
The raw data for single crystal diffractions are too large. Readers can request it from the authors via email.
Source Data Extended Data Fig. 3
Illustrated crystal structure of vesta file
Source Data Extended Data Fig. 4
Statistical source data
Source Data Extended Data Fig. 5
Statistical source data
Source Data Extended Data Fig. 6
Statistical source data
Source Data Extended Data Fig. 7
Unprocessed raw test data
Source Data Extended Data Fig. 8
Statistical source data
Source Data Extended Data Fig. 9
Statistical source data
Source Data Extended Data Fig. 10
Statistical source data. Input files and scripts for DFT calculations using VASP
Rights and permissions
Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.
About this article
Cite this article
Xie, Z., Luo, R., Ying, T. et al. Dynamic-to-static switch of hydrogen bonds induces a metal–insulator transition in an organic–inorganic superlattice. Nat. Chem. (2024). https://doi.org/10.1038/s41557-024-01566-1
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41557-024-01566-1