Material production and heat treatment
The steel ingot was continuously casted by desulphurization, converter processing, refining in furnace, and vacuum degassing. The ingots were then heat-treated by pre-heating (< 900°C), heating (950 – 1020°C) and tempering (1080 – 1120°C) for a total of 8 hours. The steel sheets were then hot-rolled in 5 steps between 1100°C and 850°C to reduce the ingot thickness from 40 mm to 6 mm, followed by water-cooling to 600°C. The steel sheets were then slow cooled in a pit furnace at 350°C.
TEM experiments
Pearlitic steel sheet samples of 10 × 10 × 1 mm were mechanical polished, followed by a dual 8-kV argon ion beam polishing at an angle of 5° for 45 minutes in a precision etching coating system (PECS, Gatan). After polishing, EBSD mapping was performed to identify suitable grains to prepare TEM samples with a zone-axis near [111] ferrite. A FIB lift-out then was performed to prepare transmission electron microscope (TEM) specimens by using a xenon ion beam in three steps: 30 kV-300 pA, 30 kV-30 pA, and 5 kV-30 pA, in a Thermofisher G4 Hydra Plasma FIB-SEM (PFIB) equipped with an EBSD detector (Symmetry, Oxford Instruments). The same procedures were used for preparing the post-mortem TEM samples from the deformed pillars. The TEM sample was imaged by a Thermo-Fisher Themis-Z Double-corrected S/TEM at 300 kV. The convergence angle is 17.9 mrad, while the collection angle under HAADF and ABF are 38 – 200 and 9 – 35 mrad, respectively. The selected-area diffraction pattern was taken with a camera length of 200 mm in a JEOL JEM-2100 TEM at 200 kV. Geometric phase analysis was conducted with a software package Strain++ (25). The ferrite lattice, far from the interface, was used as a reference to calculate the strain.
APT specimen preparation
APT samples were prepared from 1×1×15 mm matchstick bars, starting with rough electropolishing using 25% perchloric acid in acetic acid at 10-30 V until two needle specimens from the middle of the bar samples. Fine polishing was then conducted in 2% perchloric acid in butoxyethanol under an optical microscope. Polishing was undertaken at 30 V until a neck forms near the apex, and then at 10 V until a new sharp tip was formed. The sharpened tips were then ion-milled in FIB annular milling in the PFIB to capture the lamellar cementite in the APT field-of-view, by using an in-situ annular milling process as illustrated in Extended Data Fig. 1 a ~45° inclined cementite lamella was targeted as it provides an optimized combination of data yield and spatial resolution of elemental mapping at the interface. Although the analysis direction (the z-direction along the length of the tip) gives the highest spatial resolution in APT analysis26, using this analysis direction (horizontal cementite in tips) resulted in high tip fracture rate and low data yield in voltage-mode APT. The ion milling began with a 30 kV-4 nA beam for locating the region of interest (ROI) in the tip, followed by a 30 kV-100 pA beam for finer operation, ending with a 2 kV-100 pA beam for final shaping, see Extended Data Fig. 1 for details.
Hydrogen/deuterium charging and subsequent sample handling
Deuterium charging of APT tip samples was carried out by using 0.1 M NaOH in D2O for 30 seconds. The charged samples were then transferred using a cryogenic sample transfer (cryo-transfer) protocol described in the Supplementary Information of a previous work (21). The samples used for in-situ micro-compression tests were hydrogen-charged in the same charging rig by using 0.1 M NaOH in H2O for 1 hour. After a quick dusting, the hydrogen-charged samples were immediately loaded into a SEM (Ultra, Zeiss) for testing. The loading takes approximate 10 minutes. The in-situ compression test generally required 30 minutes to start and 4 hours to complete.
APT analysis
The data in Fig. 1 was acquired by using a Local Electrode Atom Probe (LEAP 3000 Si, CAMECA), and the data in Fig. 3 was acquired by using a Local Electrode Atom Probe (LEAP 4000X Si, CAMECA) with a cryo-transfer suite. All APT experiments were conducted at a pulse frequency of 200 kHz, a specimen temperature of 50 K, and a pulse fraction of 20%. All APT reconstructions used the default algorithm in AP Suite (Version 6.1, CAMECA) with the settings of 57% detector efficiency and 1.65 image compression factor. The 1-D concentration profiles in Figs. 1 and 3 were obtained from cuboid ROIs with fixed bin width of 0.1 without bin overlaps. The ion labeling is shown in Extended Data Fig. 2 from (A) deuterium-charged and (B) deuterium-free specimens. Note the absence of 2 Da peak in uncharged data (B).
Micro-compression specimen fabrication
Pearlitic micropillars were prepared as illustrated in Extended Data Fig. 3: (A) EBSD was conducted on the surface of a polished bulk specimen to identify grains with normal orientation parallel to <110>ferrite. (B) and (C) Rough annular trenching was carried out at 30 kV, 4 nA, in a pattern with an outer diameter of 15 μm and an inner diameter of 3 μm. (D) Due to the cementite-ferrite orientation relationship, lamellar cementite in the {112} planes of ferrite matrix can be obtained, (E) We selected the micropillars with uniform cementite lamellae that are 40-50 degrees inclined to the subsequent micro-compression direction, to allow interfacial slip to be the primary deformation mode. The final ion milling step used 30 kV-30 pA until the desired geometry was achieved.
Micro-compression experiments
The in-situ micro-compression experiments were conducted in a Zeiss Ultra SEM operated at 5-30 kV. Compression was applied by a PI85 (Hysitron) nanoindenter with a diamond punch with a 2-μm-diameter flat tip. Compression experiments were carried out in strain-control mode at a rate of 3 nm/s, which is equivalent to a quasi-static loading at a strain rate of 10-3. The loading and straining data and the sample geometry were used to calculate the data on the stress–strain curves. Resolved sheer stress was used to account for the influence of inclined angles of the cementite lamellae on the mechanical measurement.
Thermal desorption analyses
Thermal desorption analyses using the experimental setup detailed in referece35 were conducted to study hydrogen desorption from bulk specimens. The tests were conducted at a heating rate of 400 °C per hour on 10 × 10 × 1 mm sheets of pearlitic steel samples with cleaned and polished surfaces. This specimen dimension is the same as for the bulk specimens for the micro-compression tests. As shown in Extended Data Fig. 5, three conditions were used to confirm the hydrogen content in the three specimens: i) uncharged reference sample (black) with 17 mass part per billion (mppb); ii) 60-minute hydrogen-charging followed by a 30-minute desorption in the TDS vacuum chamber (blue) with 137 mppb; and iii) 60-minute hydrogen-charging followed by a 4-hour desorption in the TDS vacuum chamber (red) with 35 mppb. The inserted figure displays a comparison between the cases of (iii) and (i) with a clear distinction in their hydrogen content.
DFT simulations
DFT simulations were conducted by using the Vienna ab-initio simulation package (VASP)36. The simulations used the Perdew-Burke-Erzerhof (PBE) exchange-correlation functional37 and projector augmented-wave (PAW)38 pseudopotentials with Fe-3d74s1 and C-2s22p2 valence electrons and a plane wave basis set cutoff of 500 eV. The simulations included spin polarization to account for the magnetic moment of iron, and a Methessel-Paxton smearing of 0.1 eV. All structures were minimized until the energy change between three consecutive steps was below 10-5 eV. The relaxed lattice parameters of cementite (assuming the Pnma space group definition) were a = 5.03 Å, b = 6.71 Å and c = 4.48 Å, and that of ferrite was a = 2.83 Å.
Dilute carbon vacancies and hydrogen defects, and pairs of those defects, were simulated in a 192-atom supercell of cementite, built from a 2 x 2 x 3 replica of the Pnma unit cell, and a 128-atom supercell of ferrite, built from a 4 x 4 x 4 replica of the conventional bcc unit cell. K-point sampling for these supercells was through Monkhorst-Pack grids39 of 4 x 3 x 2 and 4 x 4 x 4, respectively. For higher concentrations, we describe the system using configurational ensembles, which allows us to take into account the various hydrogen-vacancy and vacancy-vacancy complexes that may form in a hypo-stoichiometric carbide. The configurational ensembles were modelled with the aid of the Site-Occupation Disorder (SOD) algorithm,40 using 64-atom 2x1x2 supercells of cementite, with a 4x5x4 k-point grid. This resulted in 128 independent configurations at the four stoichiometries considered (6.25%, 12.5%, 18.75% and 25%).
The formation energy,
, of a defect (e.g. a hydrogen interstitial atom or a carbon vacancy), was calculated as:

Where
is the DFT energy of a perfect supercell,
is the DFT energy of the same supercell with an added defect, and
is the chemical potential of any atoms added (–) or removed (+) from the perfect supercell to form the defect. For carbon vacancies,
was taken as the DFT energy of diamond (Fd
m), while for hydrogen interstitial atoms, to sidestep the limitation of DFT in modelling gaseous dimers,41,42 the chemical potential was defined with respect to a dilute H atom in ferrite:

As such, the defect formation energy for hydrogen defects is also the relative change in solution energy between cementite and ferrite,
.
The calculated carbon vacancy formation energy in otherwise stoichiometric cementite is 0.65 eV, indicating that the process is endothermic. Considering APT evidence of hypo-stoichiometry of cementite, especially near the cementite/ferrite interface, we believe that carbon vacancies exist in the cementite matrix. Thus, we compute the formation energy of vacancies at higher concentrations, shown in Extended Data Fig. 8. The lowest energy configuration for each structure is used to investigate the trapping of hydrogen.
Extended Data Fig. 9a shows the binding energy between a single vacancy and a hydrogen atom as a function of distance, modelled in the larger 192-atom supercell. Hydrogen is attracted to vacancies that are within 3.5 Å, and is preferentially accommodated within the vacancy, where it is strongly bound (binding energy of 0.11 eV). The attrition is short-ranged, with the energy rapidly converging to that of dilute hydrogen in a perfect cementite. We further consider the hydrogen trapping behavior in the structures with higher vacancy concentrations by placing a hydrogen atom in every interstitial site of the lowest energy SOD configurations for each composition. The results are shown in Extended Data Fig. 9b. A wide range of solution energies is observed, depending on the local environment of the hydrogen atom in the hypo-stoichiometric cementite. At all concentrations, all states for hydrogen solution are lower energy than in the stoichiometric cementite, suggesting that vacancies have a strong trapping effect. The presence of 6.25% vacancies leads to a reduction in solution energy of up to 0.1 eV. Interestingly, at low vacancy concentrations, hydrogen preferentially occupies the vacancy sites, while at higher vacancy concentration the interstitial sites become lower energy, with the cross-over occurring between 12.5–18.75%.
The interface of cementite is not only characterized by the presence of excess vacancies, but also by a biaxial strain field that accommodates the lattice mismatch in the coherent ferrite/cementite interface, as shown in our HAADF-STEM observations. The lattice mismatch between ferrite and cementite is 2.7% along [111]ferrite//[010]cementite and 0.8% along [
]ferrite//[10
]cementite. We consider a range of strains for both materials, varying linearly between these values and zero strain. We rotate the strain matrix to align them to the coordinates of the conventional unit cells of ferrite and cementite, and then apply them to the
ferrite supercell and the lowest solution energy structures of cementite shown in Extended Data Fig. 9b.
We define the equilibrium lattice parameter at the interface,
and
for ferrite and cementite respectively, as those for which

thus,

With the DFT-derived elastic constants of ferrite and cementite (see Extended Data Table 3), this equates to a 56% of the total strain being accommodated in ferrite, and the resulting lattice spacing on the habit plane of the interface are reported in Extended Data Table 4. The deviation in strain from this equilibrium interface biaxial strain is defined as

Extended Data Fig. 9c shows that in both materials biaxial compression increases the solution energy, while tension decreases it. In practice this means that the interfacial strain leads to a decrease in hydrogen solution energy in ferrite as H approaches the interface from bulk ferrite, and an increment in solution energy in cementite as H approaches the interface from bulk cementite. The presence of vacancies has an even stronger effect on the reduction of hydrogen solution energy, overwhelming the effect of strain in cementite.
Data and Code Availability: The data that support the findings of this study are available from the corresponding author upon reasonable request.
References for methods
35. Lin, Y. C. et al. Hydrogen trapping and desorption of dual precipitates in tempered low-carbon martensitic steel. Acta Mater.196, 516–527 (2020).
36. Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci.6, 15–50 (1996).
37. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
38. Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B50, 17953–17979 (1994).
39. Monkhorst, H. J. & Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B13, 5188 (1976).
40. Grau-Crespo, R., Hamad, S., Catlow, C. R. A. & De Leeuw, N. H. Symmetry-adapted configurational modelling of fractional site occupancy in solids. J. Phys. Condens. Matter19, (2007).
41. Burr, P. A., Murphy, S. T., Lumley, S. C., Wenman, M. R. & Grimes, R. W. Hydrogen accommodation in Zr second phase particles: Implications for H pick-up and hydriding of Zircaloy-2 and Zircaloy-4. Corros. Sci.69, 1–4 (2013).
42. Lozovoi, A. Y., Alavi, A. & Finnis, M. W. Surface energy and the early stages of oxidation of NiAl(110). Comput. Phys. Commun.137, 174–194 (2001).