Introduction

The recent observation of near-to-room-temperature superconductivity in hydrides under high pressure1,2 has demonstrated that a superconducting (SC) state with high critical temperature does not necessarily require an unconventional electron–electron coupling mechanism. The key requirements here are high phonon frequencies, such as those associated to the hydrogen (H) vibration modes, a strong coupling of these modes to the electronic states at the Fermi level, and a non-conventional structural environment (in the specific case, stabilized by pressure) that favors such a coupling. A still-open question is whether these same ingredients can be used to design new SC materials at ambient pressure—with properties of interest for applications even though at lower temperatures—or a different paradigm must be considered when designing H-rich superconductors at low pressures. This point is especially relevant in light of the recently-claimed attainment of room-temperature superconductivity in lutetium hydride at relatively low pressures3.

One of the routes to create H-rich materials is ionic liquid gating-induced protonation. As this technique makes it possible to insert H atoms in crystallographic positions unattainable by a conventional synthesis approach, it has found notable applications in oxides4,5,6,7,8,9 and demonstrated a great potential in layered materials as well10,11. Among the latter, transition metal dichalcogenides (TMDs) are particularly interesting because, despite their structural simplicity, they exhibit a variety of quantum phases12,13—charge-density wave (CDW), ferromagnetism, Mott insulating state, etc.—that may be tuned by protonation. In some cases, TMDs display a topological order, hence, once made SC, they may host Majorana fermions14,15. TMDs are also easy to handle and can be used to fabricate electronic devices16,17,18, even printed ones19,20. Hence, by tailoring their physical properties, one can realize good and affordable prototypes for various quantum-technology applications, SC electronics, or quantum computing21. In this context, the archetypal TMD compound titanium diselenide (1T-TiSe2) is an ideal candidate to benchmark the ionic-gating-induced protonation technique, since it exhibits strong electron–electron correlations and a variety of intriguing quantum phases16, akin to those found in other layered materials such as iron-based compounds22,23, cuprates24, and heavy-fermion systems25. Furthermore, the electronic ground state of 1T-TiSe2 is readily tunable by a variety of different methods, including chemical intercalation26,27,28, applied pressure29, and electrostatic gating30,31.

In this work, we show that protonation of TMDs induced by ionic liquid gating is a powerful and unique doping technique to control their electronic ground state. In particular, we demonstrate that it allows a robust and non-volatile control over CDW and SC in 1T-TiSe2. These two quantum phases dominate the phase diagram of 1T-TiSe2 and are found to coexist across all the investigated H doping levels. Fully protonated HxTiSe2 samples exhibit an impressive doping level (x ≈ 2), with negligible structural alterations, vestigial CDW signatures, and robust bulk superconductivity with possible multi-band and gapless features revealed by different experimental techniques. Our calculations indicate that these peculiarities of protonated TiSe2 are caused by the unique capability of H atoms to act not only as pure electron donors since, at sufficiently large H concentrations, their doping becomes band-selective and fills the Ti-\({d}_{{z}^{2}}\) bands, allowing the band structure of 1T-TiSe2 to mimic that of other archetypal TMD compounds, such as the SC 2H-NbSe2 or the semiconducting 2H-MoS2; furthermore, the hybridization between H orbitals and the Ti states at the Fermi level is found to be pivotal in reducing the electronic screening and thereby enhance the electron-phonon coupling. Our results establish that the role played by hydrogen in SC layered materials at low pressures is crucial and yet firmly distinct from that found in high-pressure hydrides, providing a key insight for the search of high-temperature superconductivity in layered compounds at ambient pressure.

Results

Electric field-driven hydrogen intercalation

We control the H loading in our 1T-TiSe2 samples via the ionic liquid gating-induced protonation method4,6,10,32, as depicted in Fig. 1a. When the gate voltage VG is swept in ambient conditions above a threshold voltage (Methods), the intense electric field at the TiSe2/electrolyte interface splits the water molecules absorbed in the ionic liquid and drives the H+ ions into the van der Waals gap of the TiSe2 crystal. H incorporation in the TiSe2 lattice is confirmed by nuclear magnetic resonance (NMR) and muon spin rotation (μSR) measurements as discussed later, whereas significant intercalation by the organic ions of the ionic liquid is ruled out via X-ray photoelectron spectroscopy (XPS; see Supplementary Note 1). In-situ monitoring of the room-temperature TiSe2 resistivity during gating (Methods) shows that ρ(300K) remains unaffected by the applied VG below the threshold for H intercalation, then rapidly decreases and reaches a minimum (Fig. 1b). Maintaining VG = + 3 V beyond this point leads to ρ(300 K) gradually increasing again over time, eventually reaching a plateau for gating times in excess of ~45 min and up to several days.

Fig. 1: Ionic liquid gating-induced protonation: electric transport and structural characterization.
figure 1

a Sketch of a TiSe2 crystal immersed in the electrochemical cell for ionic liquid gating-induced protonation, including the electrical connections. The side panel shows a ball-and-stick model of the HxTiSe2 structure with x 1. b Gate voltage VG and resistivity ρ(300 K) measured in situ as a function of time, during a typical gating sequence in the electrochemical cell at ambient conditions. The gating time (i.e., the time at which the crystal is kept at VG = + 3 V) is highlighted. c ρ(300 K) measured ex situ and (d), Hall carrier density nH at 4.2 K, of a series of TiSe2 crystals gated for increasing amounts of time. Error bars almost entirely arise from non-ideal sample geometry. Shaded gray bands are guides to the eye. e X-ray diffraction spectra in selected TiSe2 crystals at different gating times. The inset shows the resulting unit cell volume as a function of the gating time. f ρ as a function of temperature (T) for the same crystals reported in (c), normalized by the value at 300 K. The inset shows a magnification of the T range where the superconducting transitions are observed. g Raman spectra in ambient conditions acquired on the freshly-cleaved surfaces of a pristine TiSe2 (black curve) and a H2TiSe2 (red curve) crystal.

This gate-induced change in the electric transport properties is non-volatile. Indeed, as shown in Fig. 1c, ex-situ measurements on samples gated for a fixed amount of time and then removed from the electrochemical cell give values of ρ(300 K) that closely follow those determined in situ. The saturation of ρ(300 K) for large gating times suggests that most of the H dopants are driven into the sample in the first few tens of minutes, and further increasing the gating time does not alter the average stoichiometry appreciably. This picture is confirmed by assessing the changes to the structural and electronic properties of the protonated samples.

Low-temperature Hall effect measurements (Methods) show an increase in the Hall density nH at T = 4.2 K by more than an order of magnitude, going from ~4  1019 cm-3 in pristine TiSe2 to ~5  1020 cm-3 in H2TiSe2 (Fig. 1d), indicating that H loading introduces a strong electron doping in the protonated TiSe2 samples. At the same time, a room-temperature X-ray diffraction (XRD) analysis (Fig. 1e) shows that the H loading does not alter the 1T lattice structure of the pristine material, and simply expands the unit cell by up to ~0.9%. The lattice expansion is comparable to that driven by incorporation of Cu atoms in SC CuxTiSe2 (~0.6%)26 and much smaller than the one reported upon lithiation in SC LixTiSe2 (~12%)33. The pronounced broadening of the XRD peaks however indicates that the protonation introduces a large degree of structural disorder in the samples, which may be partially responsible for the re-entrant behavior of ρ(300 K) upon gating times in excess of those associated to the local minimum shown in Fig. 1b, c. Again, both the structural and the electronic modifications to the TiSe2 crystals occur within the first ~45 min of gating time, confirming the behavior showcased by the room-temperature resistivity.

The dramatic effect of the protonation on the electric transport properties of TiSe2 is even more evident when these are measured as a function of temperature. Fig. 1f shows the temperature dependence of the normalized resistivity ρ(T)/ρ(300 K) of TiSe2 crystals gated for different amounts of time (Methods). Pristine crystals exhibit a strongly non-monotonic ρ(T) dependence with a broad peak around ~160 K and no sign of a SC transition. This peak is typical of 1T-TiSe2 and is caused by the reconstruction of the electronic band structure upon the onset of CDW order, which shifts the Fermi level and partially gaps the Fermi surface26,27,30,34,35. The monotonically-increasing ρ(T) dependence up to TCDW is also commonly observed in high-quality 1T-TiSe2 crystals, but whether it originates from a semimetallic band overlap36,37 or from a narrow semiconducting gap in the presence of a small extrinsic electron doping38,39,40 remains currently uncertain. A small gating time of ~5 min strongly suppresses the intensity of the CDW peak and leads to the appearance of SC with an onset transition temperature \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}\simeq 3.3\) K. Larger gating times up to ~45 min lead to a further reduction in the intensity of the CDW peak and simultaneously slightly increase \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}\) to ~3.7 K (with no clear dependence, see inset to Fig. 1f). Also here, gating times in excess of ~45 min lead to a clusterization of the ρ(T) curves around a common behavior, characterized by a complete SC transition with \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}} \sim 3.7\) K and a faint trace of CDW order in the form of a slope change below ~200 K. Consistent with the XRD results, a large degree of structural disorder in these highly-doped samples is evidenced by the four-fold reduction in their residual resistivity ratio (Methods), from its maximum value ~32 attained at a gating time of ~30 min, to a final value ~8 at a gating time of 5 days.

The number of intercalated protons in HxTiSe2 samples pertaining to this state of clusterized ρ(T) and saturated nH and cell volume (see Fig. 1f, d and e respectively) is quantified by means of 1H-NMR (Methods and Supplementary Note 2). In particular, the decay of the spin-echo amplitude is measured both in a collection of fully-doped TiSe2 crystals and in a reference sample of hexamethylbenzene (C12H18). A comparison between the signal amplitudes for the two samples shows that the intercalated TiSe2 crystals contain an average number x = 2.0 ± 0.3 of protons per TiSe2 formula unit. Assuming the H concentration to be linearly proportional to nH at T = 4.2 K, this then allows estimating an average x ~ 0.4 for intercalated TiSe2 to exhibit an incomplete resistive transition with reduced \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}\) (gating time of ~5 min), and an average x ~ 1 for a complete resistive transition to develop (gating time of ~17 min).

H loading also affects the vibrational properties of TiSe2, as demonstrated by room-temperature Raman spectroscopy on freshly-cleaved surfaces (Fig. 1g; see Methods). Spectra acquired on pristine TiSe2 exhibit the two main Raman modes of 1T-TiSe2, the Eg mode at ~138 cm−1 and the A1g mode at ~202 cm−1, consistent with the literature41,42,43,44,45, and no additional peaks. In spectra acquired on H2TiSe2, the same Eg and A1g peaks are redshifted to ~136 cm−1 and ~199 cm−1 respectively. Two additional, broader peaks emerge in H2TiSe2 at ~156 cm−1 and ~260 cm−1, which could be ascribed to intercalant-activated two-phonon processes at the L and/or M points42,43,44,46, and a very broad band appears at high wavenumbers centered around ~2900 cm−1. Our first-principle calculations (see Supplementary Note 3) indicate that the appearance of the peaks at ~156 cm−1 and ~260 cm−1 can be justified by the removal of symmetry of the 1T-TiSe2 system resulting from the incorporation of H atoms in both atomic (H1TiSe2) and molecular (H2TiSe2) forms. The broad band centered around ~2900 cm−1 can instead be naturally interpreted as originating from H2 molecules confined in the TiSe2 matrix, with a renormalized phonon frequency47,48. The simultaneous observation of these multiple peaks is therefore evidence of a strong disorder in the H concentration, phase and distribution in the TiSe2 matrix. In this case, the observation of consistent Raman spectra across different spots in the H2TiSe2 crystals could be interpreted as the phase separation between regions with different H concentration to occur over length scales smaller than the laser spot size (~1 μm).

Charge density waves in HxTiSe2

The evolution of the CDW order with increasing gating time t can be more accurately tracked by considering its onset temperature, TCDW, and the intensity of the CDW peak, ICDW. As firmly established in the literature30,33,36, the former can be identified with the position of the inflection point in the ρ(T) curves, that corresponds to the minimum in dρ/dT. As shown in Fig. 2a, which displays dρ/dT as a function of T, the visibility of the local minimum is suppressed by increasing gating time, but its position remains basically unchanged and close to its pristine value TCDW ≈ 198 K. The intensity can be evaluated as ICDW = ACDW(t)/ACDW(0), where ACDW(t) is the area contained between each normalized ρ(T) curve (shown in Fig. 1f) and a line connecting its two values at 50 and 250 K33, as exemplified in Fig. 2d. The trends of TCDW and ICDW as a function of gating time are summarized in Fig. 2b (filled red hexagons and hollow black diamonds, respectively). Clearly, while TCDW is almost constant, ICDW is rapidly suppressed upon increasing gating time, and becomes of the order of 1% of its pristine value in H2TiSe2 samples. This behavior is similar to that reported in LixTiSe2 crystals33, but differs from that of CuxTiSe226,49, Ti1+xSe237, electrostatically ion-gated TiSe230, and TiSe2 under pressure29, where the suppression of ICDW was accompanied by a decrease of TCDW. The insensitivity of TCDW to increasing doping was interpreted, in the case of LixTiSe2, as a lack of full in-plane percolation of the dopants, leading to a spatial separation between SC Li-rich regions and CDW-ordered pristine regions33. The pronounced increase in structural disorder detected by both XRD and Raman spectroscopy in our samples raises the question of whether a similar picture may hold also in the case of HxTiSe2.

Fig. 2: Charge density waves in HxTiSe2.
figure 2

a First derivative of the resistivity ρ as a function of temperature T, dρ/dT, obtained by numerical derivation of the ρ(T) curves shown in Fig. 1f for increasing gating time. Red hexagons highlight the temperature TCDW at which the local minimum associated with the onset of CDW order is observed. Error bars show the uncertainty on its determination due to the experimental noise level. b Temperature-gating time phase diagram of HxTiSe2. The onset temperature of CDW order, TCDW, and the superconducting transition temperature, Tc, are plotted as a function of gating time (left scale). The CDW intensity ICDW (right scale) is calculated as ACDW(t)/ACDW(0), where ACDW(t) is the area contained between each normalized resistivity curve, ρ(T)/ρ(300 K), and a straight line connecting the data points at 50 and 250 K. Shaded areas highlight the different phases observed in the system. c T dependence of the spin-lattice relaxation rate \({T}_{1}^{\,-1}\) at μ0H 3.5 T (inset). The dashed line is a linear fit to the experimental data highlighting a well-defined Korringa-like behavior for T 170 K and a marked decrease of \({T}_{1}^{\,-1}\) upon increasing temperature above 210 K. The main panel reports the T dependence of \({\left({T}_{1}T\right)}^{-1}\), making the crossover between the different behaviors even clearer. The dashed line is a constant line corresponding to the fitting curve reported in the inset. Error bars are the standard deviation of the fit parameters in Eq. (3). d Normalized resistivity and (e), Hall coefficient RH as a function of T in three TiSe2 crystals with different ICDW. Both protonated crystals (intermediate and low ICDW) are superconducting with \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}\approx 3.7\) K, which is defined as the threshold where ρ(T) reaches 95% of its normal-state value. The shaded blue region in (d) highlights the area from which ACDW(t) is calculated.

In general, however, an explanation relying purely on phase separation between H-deficient, CDW-ordered regions and H-rich, SC regions might be too simplistic, since it is not straightforward that the presence of dopants destroys the CDW order in TiSe2 compounds. For example, scanning tunneling microscopy measurements in CuxTiSe2 indicate that CDW order robustly survives even in Cu-rich areas where the local electron doping is large50,51. On the other hand, disorder itself is able to suppress the signatures of CDW order in electric transport, as was reported in the case of irradiated NbSe252. In this context, the temperature dependence of the Hall coefficient RH (Methods and Supplementary Note 4) suggests that a weakened CDW order could survive in dopant-rich areas also in the case of HxTiSe2, as shown in Fig. 2d, e. In agreement with the established literature33,35,36,53, pristine TiSe2 exhibits both a strong CDW peak in ρ(T) (black curve in Fig. 2d, ICDW ≈ 0.9) and a change of sign in RH around T ≈ 215 K (black curve in Fig. 2e). This occurs when the onset of CDW order opens a gap at the Fermi level in part of the band structure, causing a reconstruction in the Fermi surface which changes the dominant character of the charge carriers from holonic for TTCDW to electronic for TTCDW33,35,36,53. An intermediate H doping (blue curve in Fig. 2d, e) suppresses the CDW peak in ρ(T) (ICDW ≈ 0.13) and strongly reduces the absolute value of RH, but a sign change is still clearly present at T ≈ 225 K, indicating that the Fermi surface still undergoes a reconstruction. H2TiSe2 alone (red curve in Fig. 2d, e) exhibits no sign change in RH up to ~300 K, and a CDW peak in ρ(T) reduced to a slope change (ICDW ≈ 0.016). However, the Hall coefficient remains strongly T-dependent, resulting in a tenfold reduction in nH as the sample is cooled from 300 to 4 K. This freeze-out of the free charge carriers can be ascribed to the presence of trap states introduced by structural disorder, but could also be assisted by a residual partial gapping of the Fermi surface if a weakened CDW order survives in the H-rich regions. The absence of a sign change in RH could indicate either that the CDW gap still opens around ~200 K in H2TiSe2 but is partially shifted below the Fermi level by the increasing electron doping (similar to the case of CuxTiSe250,51), or that the disorder introduced by the H intercalation suppresses long-range CDW order in the H-rich regions and allows only short-range CDW order to survive (as in the case of irradiated NbSe252). A combination of the two effects is of course also possible.

Strong indications of a reconstruction of the Fermi surface around T ~ 200 K are also obtained from NMR measurements of the 1H spin-lattice relaxation rate \({T}_{1}^{\,-1}\) in H2TiSe2 samples at μ0H 3.5 T (Methods). As shown in the inset to Fig. 2c, \({T}_{1}^{\,-1}\) exhibits a linear dependence on T for 80 K T 170 K consistent with the so-called Korringa behavior proper of relaxation processes mediated by itinerant electrons54. This is in agreement with the results of NMR measurements in other TMDs such as VSe2, VS2, and IrTe2 within the CDW phase55,56,57. Higher temperatures T 170 K mark a clear departure from the linear-in-temperature trend. In particular, a plateau of \({T}_{1}^{\,-1}\) between 170 and 210 K is followed by a suppression of \({T}_{1}^{\,-1}\) upon increasing T. The crossover between the two different behaviors is even clearer in the T dependence of \({\left({T}_{1}T\right)}^{-1}\) (main panel of Fig. 2c) and it seemingly correlates with the onset of the CDW phase. In the ideal case of a Fermi gas, it is well-known that \({\left({T}_{1}T\right)}^{-1}\propto \varrho ({E}_{F})\) – i.e., the density of states at the Fermi energy54. The observed drop of \({\left({T}_{1}T\right)}^{-1}\) for T 170 K is then strongly indicative of a progressive decrease of ϱ(EF) upon increasing temperature above the critical temperature of the CDW phase, at variance with what is observed in VSe2, VS2, and IrTe255,56,57. This peculiar behavior will be discussed in detail elsewhere58. Finally, we stress that the sharp maximum observed in the spin-lattice relaxation rate at around 10 K has already been reported in pristine TiSe259. This observation confirms that the intercalated 1H nuclei probe the intrinsic electronic properties of the host material.

Superconductivity in HxTiSe2

The evolution of the SC phase in HxTiSe2 appears completely unaffected by the presence of CDW order, as summarized by the values of Tc as a function of gating time (blue circles in Fig. 2b). A complete SC transition with a sizeable \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}} \sim 3.7\) K is observed not only in the fully-doped H2TiSe2, where no Fermi-surface reconstruction is detected by Hall effect, but also in the partially-doped HxTiSe2, where the Fermi surface is certainly partially gapped (see inset in Fig. 1f). In a magnetic field μ0H perpendicular to the ab plane of the H2TiSe2 crystal (Fig. 3a and Supplementary Note 5; see Methods) the SC transition broadens and shifts to lower T. At 2 K, a well-defined zero-resistance state (ZRS) is observed at least up to μ0H = 30 mT and the intermediate dissipative regime exhibits the typical behavior of a thermally-activated flux flow (TAFF), \(\rho (T,H)={\rho }_{0}\exp [-U(H)/{k}_{{{{{{{{\rm{B}}}}}}}}}T]\), at any T and H. This is clear from Fig. 3b, where ρ/ρ0 is plotted in logarithmic scale against T−1. The activation barrier for vortex motion, obtained from these curves, scales logarithmically with increasing magnetic field, \(U(H)={U}_{0}\ln ({B}_{0}/{\mu }_{0}H)\) (right panel in Fig. 3b), indicating a collective flux creeping60. A fit to the U(H) data gives U0 ≈ 8.9 meV and B0 ≈ 0.39 T, comparable to the values found in ion-gated TiSe231. The observation of this direct transition from ZRS to TAFF already at relatively large T 2 K is analogous to those found in overdoped LixTiSe233 and CuxTiSe226, and provides no evidence that HxTiSe2 may host the anomalous quantum metallic behavior that was claimed to destroy the ZRS at arbitrarily small magnetic fields in ion-gated TiSe231 and in under- and optimally-doped LixTiSe233.

Figure 3c shows the T-dependence of the zero-field cooled (ZFC) magnetic moment m(T) measured below ~5 K at μ0H = 1 mT (Methods). A rather sharp transition is observed at \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}=3.6\pm 0.1\) K, in reasonable agreement with the Tc value determined by transport measurements. As expected for the Meissner state, the isothermal m(H), measured at T = 2 K following a ZFC procedure, decreases linearly with increasing H, up to 1.3 ± 0.3 mT (Fig. 3d). Above this field, m(H) starts deviating from a linear dependence, thus defining the lower critical field Hc1(2 K). Upon further increasing H, m(H) first reaches a minimum at ≈ 5.8 mT and then it decreases in modulus, returning to zero at the upper critical field  100 mT, in fairly good agreement with the Hc2(T) value determined from transport measurements (where we use the criterion of 95% normal-state resistivity).

Fig. 3: Magnetotransport and magnetization of superconducting H2TiSe2.
figure 3

a, Resistivity ρ of a H2TiSe2 crystal as a function of temperature T in the superconducting transition, normalized by its normal-state value ρ0, for increasing magnetic field H applied orthogonal to the crystal ab plane. b, Same data of a plotted in semilogarithmic scale as a function of T−1, highlighting the Arrhenius behavior characteristic of thermally-activated flux flow (TAFF). The side panel shows the TAFF activation energy U as a function of μ0H. Dashed black line is the linear fit to the data in semilogarithmic scale which allows determining the characteristic energy U0 and magnetic field μ0H0. c, T-dependence of the magnetic moment m, recorded under zero-field cooling (ZFC) conditions with μ0H = 1 mT magnetic field applied perpendicular to the ab plane. The arrow indicates the T below which the magnetic moment starts deviating from its normal-state value, thus defining the diamagnetic onset temperature \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}=3.6\pm 0.1\) K. d, ZFC magnetic moment as a function of H measured isothermally at T = 2 K. The line is a guide to the eyes. The inset shows a magnification of the low-field region, with a linear fit on the first six points (line). The arrow indicates the field above which the magnetic moment starts deviating from a linear dependence, thus defining the first critical field Hc1 = 1.3 ± 0.3 mT. e, H − T phase diagram of H2TiSe2 constructed by combining the results from resistivity and magnetization measurements on multiple different crystals. Shaded blue regions highlight the Meissner state and the TAFF state. Red and indigo circles are obtained from the data shown in a and d respectively. Blue and green circles are obtained from Supplementary Fig. 6a and b respectively. Dashed blue lines are linear fits to the resistive critical magnetic field Hc2 (defined as the threshold for reaching ρ(T) = 0.95ρ0) above and below the kink at T/Tc ≈ 0.85. The resistive Hc2 data for CuxTiSe226 and ion-gated TiSe230 are shown for comparison.

By combining the resistivity- and magnetization measurement results, we can construct the SC TH phase diagram of H2TiSe2 shown in Fig. 3e. The Meissner state exists for magnetic fields HHc1, above which SC vortices penetrate the sample volume and move in the TAFF regime until H = Hc2, here defined as the field where ρ reaches 95% of its normal-state value. The T-dependence of Hc2 shows a prominent kink around ~0.85Tc, where its slope changes abruptly from ≈ − 0.23 to ≈ − 0.13 T/K. These two slopes correspond to an in-plane coherence length ξab(0) of 26 and 20 nm, respectively (Methods). We note that, as shown in Fig. 3e, this behavior is invariably observed across all the measured H2TiSe2 samples with little sample-to-sample variation, even in those that exhibit a strongly biphasic resistive transition (Supplementary Fig. 6a) or a weakly-localized normal state (Supplementary Fig. 6b). This indicates that it is mostly insensitive to disorder and sample inhomogeneities, and therefore is likely an intrinsic property of the material. As a consequence, since this feature deviates considerably from the single-band mean-field behavior exhibited by both the 3D CuxTiSe2 and the ion-gated 2D TiSe2, it suggests that HxTiSe2 may be a multi-band superconductor, where it is instead commonly observed61,62,63.

Further insight into the SC properties of the H2TiSe2 is obtained by transverse-field (TF) μSR measurements64,65 (Methods and Supplementary Note 6). Figure 4a, b show the zero-field cooling (ZFC) TF-polarization curves recorded either in an applied field parallel to the ab-planes, μ0H = 14 mT, or parallel to the c-axis, μ0H = 10 mT. As to the former (panel a), no significant differences were detected between datasets taken at T = 0.28 K and T > Tc. For fields applied parallel to the c-axis, instead, clearly distinct datasets are observed in the SC- and the normal state. The volume fraction of those muons probing the vortex state, fsc, was determined by fitting the time-dependent TF polarization data by using a two-step calibration procedure [see details in Methods and Eq. (4)]. We find a lower limit for the SC volume fraction fsc = 44 ± 1%, thus confirming the bulk nature of HxTiSe2 superconductivity, leaving however open the possibility that some sample regions may still be normal, as expected in case of an inhomogeneous distribution of H dopants.

Fig. 4: Muon spin rotation in superconducting H2TiSe2.
figure 4

a, b Zero-field cooling transverse-field muon-spin polarization PTF(t) in a 14-mT magnetic field applied parallel to the ab planes, (a), or in a 10-mT field applied parallel to the c-axis, (b). Solid lines are fits to the model described by Eq. (4). The error of the positron count for each bin is its standard deviation, while the error bars of each bin in PTF(t) are calculated by error propagation. c Temperature dependence of the depolarization rate (red circles) in a magnetic field μ0H = 10 mT, applied parallel to the c axis. Note the lack of saturation at low temperatures. The dark-blue square represents the depolarization rate recorded after field cooling in the same applied magnetic field. The gray dashed line represents the average nuclear dipolar contribution, here, σn = 0.32 ± 0.02 μs−1. The error bars are the standard deviation of the fit parameters.

In the ZFC case, the temperature dependence of the in-plane Gaussian depolarization rate of the implanted muons σ(T), resulting from fits to Eq. (4) (Methods), is shown in Fig. 4c. It includes the contributions of σn and σsc, the depolarization rates in the normal and SC state, respectively, and is given by \(\sigma =\sqrt{{\sigma }_{{{{{{{{\rm{n}}}}}}}}}^{2}+{\sigma }_{{{{{{{{\rm{sc}}}}}}}}}^{2}}\). σn is due to the contribution of the randomly-oriented nuclear dipole moments, here fully dominated by the intercalated H nuclei, whose moments are only partially quenched by the applied magnetic field. Therefore, the observation of a finite σn = 0.32 ± 0.02 μs−1in the normal state (dashed line in Fig. 4c) provides an independent confirmation of the sizeable H intake in the H2TiSe2 samples. The effective magnetic penetration depth was estimated by measuring also a TF spectrum at base T (dark-blue square in Fig. 4c), now using a standard field-cooling (FC) procedure, which provides a SC-depolarization rate σsc = 0.14 ± 0.03 μs−1 at T = 0.28 K. This, in turn, implies a rather large value for the in-plane magnetic penetration depth, λab = 900 ± 100 nm (Methods), indicative of a low superfluid density. Furthermore, the T dependence of σ does not show any signs of saturation down to the lowest temperature (0.28 K). At the same time, auxiliary zero-field measurements confirm that time-reversal symmetry is preserved in H2TiSe2 (see Supplementary Note 7). This result suggests that H2TiSe2 hosts a gapless-like SC state, thus making future spectroscopic investigations of the symmetry of SC gap(s) of this material highly desirable.

Density functional theory of HxTiSe2

The effects of H intercalation on the structural, electronic and SC phase of 1T-HxTiSe2 were investigated from a theoretical standpoint by means of ab initio density functional theory (DFT) calculations (Methods and Supplementary Note 8). Simulations based on a random-search algorithm reveal the existence of several metastable intercalation sites for H in the 1T-TiSe2 lattice, indicating that it can easily be trapped (at T = 0) in different local energy minima. In particular, for concentrations up to x  1 H preferentially sits in the van der Waals gap at the bridge position between two Se atoms belonging to different planes (Fig. 5a).

Fig. 5: First-principles electronic and phononic structures of HxTiSe2.
figure 5

a Ball-and-stick model of the protonated 1T-TiSe2 structure in the ground-state adsorption site with one H atom placed in the van der Waals gap at the Se-Se bridge position (also shown in Fig. 1a). The Brillouin zone with high symmetry points is also shown. b Electronic band structure and density of states of 1T-H1TiSe2 in the structure shown in panel (a) (red) compared with the pristine 1T-TiSe2 dispersion (gray). c Orbital character of the electronic band structure of 1T-H1TiSe2, with the \({d}_{{z}^{2}}\) character of the bands indicated by the size of the black circles. d Phonon dispersion relations (left panel), density of states (central panel), and electron-phonon spectral function α2F(ω) (right panel) of pressurized 1T-H1TiSe2. The spectral contributions of the H-derived modes are highlighted as red dashed lines. The right panel also reports the frequency-dependence of the total electron-phonon coupling λ(ω) as a solid blue line.

From the electronic standpoint, for x  0.1 H acts as an electron donor, rigidly shifting the Fermi level in the conduction band of TiSe2 where the electronic states exhibit a predominant in-plane character (Supplementary Fig. 10a). The simple rigid-band-shift picture breaks down at intermediate concentrations (x ≈ 1), where H doping leads to a strong deformation of the intrinsic electronic dispersion of 1T-TiSe2 (shown as the gray curve in Fig. 5b), induces a sizeable hybridization between the TiSe2 bands and the H-derived orbitals, and completely switches the orbital character of the bands which cross the Fermi level (Fig. 5b, red curve). Indeed, the positive potential induced by the ionized H atoms in the van der Waals gap of 1T-TiSe2 lowers the energy of the bands with orbital character extending in the inter-layer gap, like the \({d}_{{z}^{2}}\) band (the intensity of which is depicted as solid circles in Fig. 5c). This band structure engineering is crucial for the development of SC order, since the \({d}_{{z}^{2}}\) character of the states at the Fermi energy was already found in the SC TMD compounds NbS2 and NbSe2, where the \({d}_{{z}^{2}}\) bands are partially-filled66,67,68.

Conversely, a further increase in doping level, to the concentration determined in fully-doped HxTiSe2 samples by NMR (x ≈ 2), leads to the emergence of two possible phases. One is characterized by the formation of H2 molecules in the van der Waals gap and a band structure akin to that of intrinsic 1T-TiSe2 (Supplementary Fig. 10b). The other is a metastable insulating phase, achieved when the stoichiometry is exactly 2 H atoms per unit cell (Supplementary Fig. 10d). The actual HxTiSe2 samples are likely to be characterized by a strong disorder in the distribution of H atoms, and will thus exhibit a mixture of different single phases which however all turn out to be dynamically unstable (see Supplementary Note 8 for further details). This in turn prevents a full ab-initio calculation of the properties of the real system, in particular for the calculation of the SC critical temperature.

To have solid first-principles predictions of the SC phase even in the high density regime, the metallic 1T-H1TiSe2 phase was therefore artificially stabilized by slightly reducing the lattice constants by 4% (that corresponds to an external pressure of ~10 GPa, see Methods). This approach, already exploited to remove the CDW distortion in pure 1T-TiSe269 and reduce anharmonic effects in H-doped palladium alloys70, allows obtaining a dynamically stable high-doping phase while at the same time leaving the main features of the electronic properties of 1T-H1TiSe2 mostly unaffected (as shown in Supplementary Fig. 11). The resulting phonon band structure shows the appearance of H-derived branches (highlighted in red in the phonon density of states shown in Fig. 5d), comprising a high-energy band around 1300 cm−1 and more entangled branches at lower frequencies, close to the TiSe2 modes around 500 cm−1. Overall, the pressurized H1TiSe2 phase is therefore able to mimic all the key modifications introduced by the H dopants in the real TiSe2 samples.

In particular, the dynamical stability of the pressurized H1TiSe2 structure allows computing the electron-phonon spectral function α2F(ω), shown in Fig. 5d, from which both the electron-phonon coupling λ and the SC Tc can be determined (Methods). In sharp contrast with the high-pressure hydrides32,71, the H-derived phonon branches do not sensibly contribute to the total coupling λ = 1.15, which is however found to be significantly higher with respect to that of the low-doping structure (λ = 0.5, see Supplementary Note 8). As a consequence, H1TiSe2 is predicted to become SC at Tc ~ 13 K (Methods), a critical temperature surely higher than the experimentally-measured one, but of the same order of those measured in other SC TMDs15,67,68. This overestimation can be ascribed to differences between the real material and the computational model, with disorder effects and doping inhomogeneities easily lowering the real Tc with respect to the ideal case considered in the calculations. Anyway, the much larger predicted Tc with respect to that of the low-doping structure (Tc ~ 1.5 K, see Supplementary Note 7) highlights the crucial role played by the H dopants in determining the properties of the SC phase. Note that this enhancement cannot be ascribed merely to the application of pressure to the TiSe2 structure, since the Tc of undoped 1T-TiSe2 at 10 GPa is  1 K29,69.

Discussion

Our results show that ionic liquid gating-induced protonation allows for a robust and non-volatile tuning of the electronic ground state of archetypal correlated layered compound 1T-TiSe2. We demonstrate that this tuning is qualitatively different from what is obtained by other doping methods, such as electrostatic gating and Cu or Li intercalation. An increased level of hydrogen doping suppresses the intensity of the CDW phase of undoped TiSe2 and triggers the onset of superconductivity. Still, vestigial traces of the CDW ordering, observed even in H2TiSe2—a stoichiometry unattainable via Li or Cu intercalation—hint at the coexistence of both phases in a wide doping range. This coexistence is likely enabled by the strongly disordered nature of the hydrogen-rich compound, which hosts a mixture of different phases with distinct electronic structures determined by the several possible hydrogen intercalation configurations. The observed SC phase, possibly gapless and multi-band in character, together with the absence of an anomalous metallic phase are also at odds with what is reported in other SC TiSe2 compounds. Our experimental results are consistent with ab initio calculations showing how the effect of H intercalation is not limited to a rigid-band doping, but can lead to a full band-structure engineering of the undoped TiSe2 compound. The role of hydrogen is crucial in determining the emergence of superconductivity through different mechanisms: (i) at low concentrations, H intercalation rigidly charge-dopes the system, weakening the CDW ordering and increasing the electron–phonon coupling at the Fermi level; (ii) at higher concentrations, the doping mechanism changes, triggering a band inversion that cannot be described within a rigid band shift; this leads to the partial filling of the highly-coupled \({d}_{{z}^{2}}\) band, mimicking the exact band population found in Nb-based transition-metal dichalcogenides of the 2H polytype, such as 2H-NbSe2 and 2H-NbS2, which are superconductors even without doping; (iii) H intercalation also induces the hybridization of these \({d}_{{z}^{2}}\) states with H-derived orbitals, reducing the screening of the phonon perturbations and thereby further increasing the electron-phonon coupling. This clearly indicates that the SC phase realized in HxTiSe2 should be of a different nature with respect to those observed in CuxTiSe2 or in electron-doped MoS2, reflecting the different orbitals involved in the electron–electron pairing and the capability of H doping to hybridize the electronic states at the Fermi level and reduce the screening to phonon perturbations. At the same time, its origin is firmly distinct from that found in high-pressure hydrides, since in HxTiSe2 the high-frequency phonon spectral contributions due to the light H atoms remain weakly coupled to the electronic states at the Fermi level.

In summary, our findings show ionic liquid gating-induced protonation to be a unique doping technique, distinct from the more ubiquitous chemical substitution and alkali intercalation. Furthermore, the understanding of the difference between the mechanisms at play in hydrogen-rich superconductors at megabar pressures71 and in HxTiSe2 will be crucial in the quest for new high-Tc superconductors at ambient and low pressures, as also suggested by the room-temperature superconductivity recently claimed to appear in lutetium hydride at relatively low pressures3. Gate-driven protonation will enable accessing new electronic phases of the host compounds at ambient pressure, where hydrogen not only injects high-frequency phonon spectral contributions, but can play a triple role as a charge dopant, a source of band-selective filling, and a knob to tune the electronic screening via orbital hybridization.

Theoretical proposals in hydrogen-doped systems are recently appearing, confirming our experimental and theoretical analysis. All studies highlight how superconductivity can be induced in different materials by hydrogen incorporation. Among others, we cite the hydrogen-induced SC phases in PdCu70, diborides like MgB272 and TiB273, chromium74, Mo2C375, hBN76, and in carbon nanostructures77. The experimental realization of these theoretical proposals, made possible by the simple and reliable technique shown in this work, can reverse the accepted paradigm to access high-temperature SC phases of hydrides from the high-pressure side. Here we propose an alternative route for stabilizing new hydrogen-rich compounds, starting from promising materials at ambient pressure, driving them to the predicted SC state and then, eventually, further enhancing their SC properties by applying low or moderate pressures.

Methods

Ionic liquid gating-induced protonation

Freshly-cleaved 1T-TiSe2 crystals (HQ Graphene; typical size 1.0 × 0.5 × 0.05 mm3) were electrically contacted by drop-casting small droplets of silver paste (RS Components) to thin gold wires and immersed in a Duran crucible (40 mm diameter) filled with 1-ethyl-3-methylimidazolium tetrafluoroborate ionic liquid (EMIM-BF4, Sigma Aldrich) together with a platinum (Pt) counter electrode. The gate voltage VG was applied to the Pt electrode at 300 K in ambient atmosphere by an Agilent B2961 power source. The applied VG was always limited to a maximum value of +3 V and the gating temperature to 300 K so as to avoid undesired electrochemical reactions between the sample and the ionic liquid such as sample etching78 or organic-ion intercalation79. The four-wire resistance R = Vxx/IDS was monitored in situ by sourcing a constant current IDS ≈ 100 μA between the outer drain (D) and source (S) contacts with a Keithley 220 current source, and measuring the longitudinal voltage drop Vxx between the inner voltage contacts with an HP3457 multimeter. The resistivity was then determined as ρ = Rtwl−1, where t and w are the sample thickness and width, and l is the distance between the inner voltage contacts. Common-mode offsets were removed using the current-reversal method. The protonated TiSe2 samples were then extracted from the cell and rinsed with acetone and ethanol to remove ionic-liquid residues before further ex-situ characterizations. In between measurements, samples were stored in standard desiccators either under low vacuum or in argon atmosphere to avoid moisture contamination.

Electric transport measurements

All protonated TiSe2 crystals discussed in this work were characterized via ex-situ temperature-dependent resistivity measurements carried out in the high-vacuum chamber of a Cryomech pulse-tube cryocooler with a base temperature of ≈ 2.8 K. The magnetotransport properties of selected crystals were measured either up to ≈ 4.5 K in the variable-temperature insert of a 4He Oxford cryostat, or up to ≈ 300 K in a Quantum Design physical properties measurement system (PPMS), both equipped with 9 T SC magnets. The resistivity was determined as in the room-temperature gating runs, except that the source-drain current was sourced via an Agilent B2912 source-measure unit (Oxford cryostat) or using a standard 4-probe technique with the Resistivity option for the PPMS. The user bridge board of the PPMS has a digital-to-analog converter (DAC) which adjusts the excitation current and a delta-sigma A/D converter which reads the voltage output. In AC mode, the user bridge board applies a DC excitation to the sample and at each measurement it reverses the current averaging the absolute value of the positive and negative voltage readings. This operation eliminates errors from DC offset voltages and produces the most accurate readings. Both the longitudinal (Vxx) and transverse (Vxy) voltage drops were measured via an Agilent 34420 nanovoltmeter (Oxford cryostat) or by means of a standard 6-terminal method with the Resistivity option for the PPMS. The magnetic field H was always applied orthogonal to the samples’ ab plane.

The Hall coefficient was determined from the antisymmetrized Rxy = Vxy/IDS data as:

$${R}_{{{{{{{{\rm{H}}}}}}}}}=\frac{t}{{\mu }_{0}}\times \frac{d{R}_{{{{{{{{\rm{xy}}}}}}}}}(T)}{dH}$$
(1)

and the Hall density was calculated as \({n}_{{{{{{{{\rm{H}}}}}}}}}={(e{R}_{{{{{{{{\rm{H}}}}}}}}})}^{-1}\) where e is the elementary charge.

In SC samples, the residual resistivity ratio was determined as ρ(300 K)/ρ0, where ρ0 is the resistivity value in the normal state immediately above the SC transition; the onset transition temperature \({T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}\) was determined as the threshold where ρ(T) reaches 95% of ρ0; and the zero-temperature in-plane coherence length was determined as:

$${\xi }_{{{{{{{{\rm{ab}}}}}}}}}(0)=\sqrt{\frac{{\Phi }_{0}/2\pi }{-{\mu }_{0}(d{H}_{{{{{{{{\rm{c2}}}}}}}}}/dT)\,{T}_{{{{{{{{\rm{c}}}}}}}}}^{{{{{{{{\rm{on}}}}}}}}}}}$$
(2)

where Hc2 is the magnetic field at which ρ(H) reaches 95% of its normal-state value, μ0 is the vacuum permeability, dHc2/dT is the slope of the Hc2 − T curve, and Φ0 is the magnetic flux quantum.

X-ray diffraction

X-ray powder diffraction patterns of the TiSe2 samples were collected ex situ on selected single crystals after thorough removal of the ionic liquid residues and of the electrical leads using a 114.6 mm Gandolfi camera, with Ni-filtered Cu K-α radiation source and an exposure time of 48 h. Diffraction was impressed on a photographic film and reduced using the software X-RAY80 to obtain the intensity profiles as a function of the diffraction angle.

Raman spectroscopy

Raman spectra were acquired on freshly-cleaved surfaces in ambient conditions using a Renishaw InVia H43662 micro-Raman spectrometer. All spectra were acquired using an excitation wavelength of 514 nm, a laser power <1 mW focused through a 100X objective, an exposure time of 20 s, and 50 accumulations.

Nuclear magnetic resonance

1H-NMR measurements at μ0H 3.5 T (magnetic field parallel to the crystallographic ab plane) were performed using a TecMag Apollo spectrometer coupled to a resonant circuit made of an in-series combination of a 470 pF capacitor and of a seven-loop solenoidal coil. The coil was deformed in order to mimic the flake-like shape of the crystals and to maximize the geometrical filling factor, in turn. The spin-lattice relaxation time T1 was quantified via a conventional inversion-recovery pulsed sequence. In particular, \({T}_{1}^{\,-1}\) was extracted based on a best-fitting of the experimental recovery curves based on a stretched-exponential function

$$M(\tau )=M(\infty )\left\{1-2f\exp \left[-{\left(\frac{\tau }{{T}_{1}}\right)}^{\beta }\right]\right\}.$$
(3)

Here, M(τ) is the component of the nuclear magnetization along the quantization axis at time τ after the first inversion radio-frequency pulse, M() is the equilibrium magnetization, f ≤ 1 allows for non-ideal inversion conditions and β is the stretching exponent. β ~ 0.9 was quantified independently on temperature, in good agreement with the purely-exponential behavior expected for spin-1/2 nuclei.

Magnetization measurements

Systematic measurements of the magnetic moment as a function of temperature and magnetic field were carried out on a small H2TiSe2 single crystal with approximate surface ~1 mm2 by means of a SC quantum interference device (SQUID) magnetometer. Since the sample mass was less than 0.1 mg (i.e., below the sensibility threshold of a standard laboratory balance), the magnetic field had to be applied perpendicular to the sample surface (i.e., to ab planes). Such setup maximizes the effects of the demagnetizing factor, thus enhancing the diamagnetic signal in the SC state. The unavailability of a reliable value for the sample mass (or volume) prevents us from providing an accurate estimate of the shielding fraction in the SC state.

Muon-spin rotation measurements

Muon-spin rotation (μSR) is an extremely sensitive probe of the local (i.e., microscopic) electronic properties, which uses spin-polarized positive muons implanted in the sample under test64,65. Depending on the material density, muons typically penetrate over a depth of several hundreds of microns and, thus, are implanted homogeneously over the whole sample volume. Because of this, muons are considered as a bulk probe of matter. All the μSR measurements were carried out at the Dolly spectrometer (πE1 beamline) of the Swiss Muon Source at the Paul Scherrer Institute, Villigen, Switzerland. A 3He cryostat was used to reach temperatures down to 0.28 K. A mosaic of H2TiSe2 single crystals was glued using GE-varnish on a 25-μm thick copper foil, here acting as a thermal link (Supplementary Fig. 7). A 100-μm-thick high-purity silver degrader was fixed in front of the sample. All measurements were performed using an active veto scheme, thus removing the background signal due to muons which miss the sample. μSR measurements were performed following both standard FC and nonstandard ZFC protocols. In the former case, a regular lattice of vortex lines is established over the sample volume. This is a key requirement in order to accurately determine the magnetic penetration depth and the superfluid density. In the ZFC case, instead, we induce an artificially disordered vortex lattice in the SC state that, in turn, causes highly inhomogeneous local magnetic fields at the implanted muon sites and, hence, a much higher muon-spin depolarization rate than that observed in a conventional FC- (or normal-state) experiment. This procedure is particularly suited for evaluating the SC volume fraction of superconductors with a weak magnetic-flux expulsion, as the H-doped TiSe2 in the present case.

Muon spin rotation fits

In all cases, the time-dependent TF polarization was fitted using the following function:

$${P}_{{{{{{{{\rm{TF}}}}}}}}}(t) = \, {f}_{{{{{{{{\rm{sc}}}}}}}}}\cos ({\gamma }_{\mu }{B}_{\mu }t+\phi )\exp \left(-{\sigma }^{2}{t}^{2}/2\right)\\ + {f}_{{{{{{{{\rm{tail}}}}}}}}}\cos ({\gamma }_{\mu }{B}_{{{{{{{{\rm{tail}}}}}}}}}t+\phi )\exp (-\Lambda t)$$
(4)

Here, ftail = 1 − fsc is the fraction of muons implanted in the silver degrader and in the non SC parts of the sample, γμ/2π = 135.53 MHz/T is the muon gyromagnetic ratio, Bμ and Btail are the magnetic fields probed by the muons implanted in the sample and in the silver degrader, and ϕ is a common initial phase. Finally, Λ is the relaxation rate originating from the muons implanted in the silver degrader.

Both Λ and ftail parameters were obtained in a two-step calibration via a ZFC TF experiment at 0.28 K (see Fig. 4). First, the long-time tail (t > 5 μs) was fitted with a single oscillating component (fsc = 0), determining Λ = 30 ± 6 ns−1. In this case, due to the fast signal decay in the SC phase, the long-time tail reflects only those muons implanted in the silver degrader (and in non-SC parts of the sample). Secondly, by fixing Λ to the previously determined value, the data were fitted over the whole time range to obtain ftail = 0.56 ± 0.02. These two parameters were then kept fixed during all the subsequent TF-fits.

From the SC depolarization rate σsc, under the assumption of an ideal triangular vortex lattice, and by considering that the applied field is negligibly small with respect to the upper critical field determined from resistivity and magnetization measurements (Hc2 > 0.1 T), the magnetic penetration depth is determined as \(\lambda ={(0.00371{\Phi }_{0}{\gamma }_{\mu }^{2}/{\sigma }_{{{{{{{{\rm{sc}}}}}}}}}^{2})}^{1/4}\)81,82.

Density functional theory calculations

First-principles calculations were performed using the QUANTUM ESPRESSO package83,84. For all calculations, the experimental lattice parameters and the Generalized Gradient Approximation (GGA) for the exchange and correlation energy were adopted. This is because in 1T-TiSe2 the GGA functional, on top of the experimentally determined lattice constants, successfully predicts both the internal structural parameters (Se-Ti distance) and the dynamical instabilities guiding the system to the 2 × 2 × 2 charge density wave phase85, in perfect agreement with experiments. Conversely, since the GGA functional does not describe perfectly the electronic properties of TiSe2 due to the strong correlation effects arising from the localized d orbitals of Ti85,86, the effects of local correlations were accounted for by including a Hubbard-like correction for Ti-d orbitals in a GGA+U approach87. The U parameter, determined from first principles85,88, was set to U = 3.9 eV, allowing to reproduce the Fermi surface experimentally determined via angle-resolved photoemission spectroscopy86,89,90. Electron–ion interaction was described with pseudopotentials which include the 3s, 3p, 4s and 3d valence states for Ti and 4s, 4p and 4d for Se, with an energy cutoff up to 70 Ry (840 Ry for charge density). The integration over the Brillouin Zone (BZ) was performed using a uniform 24 × 24 × 14 grid (for the TiSe2 unit cell, and properly rescaled for the supercell calculations) with a 0.005 Ry Gaussian smearing. The electronic density of states are evaluated sampling the BZ on a 30 × 30 × 18 uniform grid using tetrahedron method for integration91. Several structural models, starting from low concentration regime up to higher doping, were considered using a random-search algorithm (over 100 configurations) to determine the H atom’s lowest-energy intercalation site. One H atom is placed in a 2 × 2 × 2 1T-TiSe2 supercell in the lowest concentration (x = 0.125), whereas two H atoms are included in the 1 × 1 × 1 unit cell (x = 2) in the highest. In the low concentration regime, we unfold the electronic bands onto the 1T-TiSe2 BZ by means of the Bands-UP software92,93. The application of an external pressure of ~10 GPa was realized by reducing the lattice constants of about 4% at constant c/a ratio, where c and a are the out-of-plane and in-plane lattice constants respectively. Phonon calculations were performed using the Density Functional Perturbation Theory94 in the harmonic approximation sampling the reciprocal space with a 4 × 4 × 4 grid. The electron–phonon coupling was converged using a 48 × 48 × 28 k-point grid for the electronic momenta. The SC critical temperature was estimated using the Allen-Dynes modified Mc-Millan equation95,96 with μ* = 0.169,97 (see also Supplementary Note 8).