The first magnetic 2D material discovered, monolayer (ML) CrI3, is particularly fascinating due to its ground state ferromagnetism. However, because ML materials are difficult to probe experimentally, much remains unresolved about ML CrI3’s structural, electronic, and magnetic properties. Here, we leverage Density Functional Theory (DFT) and high-accuracy Diffusion Monte Carlo (DMC) simulations to predict lattice parameters, magnetic moments, and spin–phonon and spin–lattice coupling of ML CrI3. We exploit a recently developed surrogate Hessian DMC line search technique to determine CrI3’s ML geometry with DMC accuracy, yielding lattice parameters in good agreement with recently published STM measurements—an accomplishment given the ∼10% variability in previous DFT-derived estimates depending upon the functional. Strikingly, we find that previous DFT predictions of ML CrI3’s magnetic spin moments are correct on average across a unit cell but miss critical local spatial fluctuations in the spin density revealed by more accurate DMC. DMC predicts that magnetic moments in ML CrI3 are 3.62 μB per chromium and −0.145 μB per iodine, both larger than previous DFT predictions. The large disparate moments together with the large spin–orbit coupling of CrI3’s I-p orbital suggest a ligand superexchange-dominated magnetic anisotropy in ML CrI3, corroborating recent observations of magnons in its 2D limit. We also find that ML CrI3 exhibits a substantial spin–phonon coupling of ∼3.32 cm−1. Our work, thus, establishes many of ML CrI3’s key properties, while also continuing to demonstrate the pivotal role that DMC can assume in the study of magnetic and other 2D materials.

2D materials represent an exciting new frontier for materials research.1 Due to their reduced dimensionality, these materials tend to exhibit stronger and longer-range electron correlation than their 3D counterparts that give rise to exotic new physics and phase behavior, including Moiré patterns,2,3 two-dimensional superconductivity,4 and exotic spin and charge density waves.5–7 The properties of 2D materials can also be tuned just by stacking,8–10 crinkling,11 straining,12–14 and twisting15–18 them. This versatility facilitates the layer-by-layer construction of designer materials with unique properties through the careful selection and ordering of their constituent layers.19 

An exciting recent development in this regard is the discovery of new magnetic 2D materials.20 While the Mermin–Wagner theorem21,22 prohibits finite-temperature magnetism for the isotropic Heisenberg model in 2D, magnetic 2D materials such as CrI3,23 Fe3GeTe2,24 and VSe225 have large anisotropies arising from strong spin–orbit couplings (SOCs) or structural anisotropies that break continuous symmetries and allow for their finite-temperature long-range magnetic order. 2D magnetic materials are of great interest as they may accelerate the development of next-generation spintronic data transmission26 and storage27 technologies. Furthermore, if integrated into 2D heterostructures, magnetic 2D materials may additionally be able to induce magnetism in nearby nonmagnetic layers through proximity effects, enabling the realization of magnetic graphene.28 Few-layer magnetic materials that have particularly strong magneto-optical responses and/or electron–phonon coupling, e.g., MPSe3/TMD heterointerfaces,29 with M = Mn, Fe, WSe2/CrI3 heterostructures,30 or bilayer (BL) CrI3,31 are also potentially promising for applications that bridge spintronics, photonics, and phononics.

However, in order to realize the potentials of 2D materials, their structural, electronic, magnetic, and optical properties have to be well characterized and understood. Experimentally, this is very challenging for mono- and few-layer systems. Monolayers (MLs) are often grown on substrates that can distort their intrinsic geometries.32–34 Due to their inherently small thickness, monolayers are also not immediately accessible to neutron and other scattering experiments, which typically require a critical thickness to yield meaningful scattering patterns.35 Given these experimental challenges, it is imperative to predictively and accurately model these materials’ atomic geometries, spin densities, and magnetic moments using first principles methods in order to elucidate the origins of their magnetic anisotropy and also to inform models upon which more macroscopic predictions can be built.36,37 Recent advances in optical response-based methods, such as fluorescence and magneto-optical Kerr effect (MOKE) microscopy, can resolve magnetic moments down to sub-micrometer scale,37 and single-spin microscopy can improve this resolution down to nanometer scale,38 complementing such first principles-based predictions.

To date, the overwhelming majority of first principles modeling of these materials has relied on Density Functional Theory (DFT).39 Even though DFT has led to transformative changes in our understanding of materials and chemistry and typically yields good results for ground state properties in 3D, especially lattice constants, it can yield widely varying results for the properties of 2D materials, such as lattice parameters and bandgaps, depending upon the functional employed. Recent studies of 2D materials have shown that DFT consistently overestimates interlayer binding energies between 2D monolayers40 and predicts lattice parameters that routinely differ from experimental ones by several percent.41,42 In contrast, Diffusion Monte Carlo (DMC), a real-space, many-body quantum Monte Carlo (QMC) method, routinely predicts lattice constants to within 1% and bandgaps to within 10% of their experimental values for the same materials.41–43 It is worth emphasizing that DMC obtains such high-accuracy results for all properties within a single, consistent framework with few approximations; moreover, the approximations can be systematically improved upon because of the variational nature of DMC. In contrast, obtaining accurate DFT results of different properties often necessitates using a different functional for each and every property, which substantially reduces the predictive power of such calculations.

In this work, we use a combination of DFT and DMC44 methods to construct a full atomistic picture of the physics and properties of the CrI3 monolayer (ML). ML CrI3 is a strongly correlated, magnetically tunable Mott insulator with a hexagonal lattice structure in which each Cr atom is coordinated by six I atoms to form a distorted octahedron (see Fig. 1).45 Atomic force microscopy and scanning tunneling microscopy point to a lattice constant of roughly 7 Å32 and a monolayer thickness of 0.7 nm.23,32 What makes ML CrI3 particularly intriguing as a material is its magnetism. In its monolayer form, CrI3 has been demonstrated by MOKE microscopy23 and single-spin nitrogen vacancy microscopy38 experiments to be a ferromagnet. The strong magneto-crystalline anisotropy with an out-of-plane easy axis makes the Ising model a natural starting point for modeling the magnetic properties of CrI3 as it allows for finite-temperature long-range order in 2D.21 However, several DFT studies have suggested that the material may be better described by the Heisenberg and Kitaev models.46,47 These studies concluded that the ligand superexchange within slightly distorted Cr–I crystal environments46 and spin–orbit coupling47 ultimately stabilize ML CrI3’s ferromagnetism. However, one DFT study of bilayer (BL) CrI3 found non-negligible magnetic moments present on the iodine atoms,48 suggesting that a more comprehensive magnetic model may be needed to quantify all of the relevant exchange interactions. In addition to its magnetism, CrI3 exhibits strong electron–phonon coupling and magnetoelasticity, as evidenced by its alternating ferromagnetic (FM) odd-layered structures/antiferromagnetic (AFM) even-layered structures9 and magnetic ordering-dependent Raman spectra.49 Although computational studies of ML CrI3’s properties have previously been performed using a variety of DFT functionals, including the generalized gradient approximation (GGA),45,50,51 hybrid functionals such as Heyd–Scuseria–Ernzerhof (HSE06),51 and the Local Density Approximation (LDA) with added spin–orbit coupling (SOC) corrections (LDASOC),52 they have yielded conflicting information regarding magnetic moments, lattice constants, and exchange parameters because of the strong functional dependence of these properties (see Table SI of the supplementary material). These discrepancies underscore the need to employ many-body approaches beyond DFT with fewer or no uncontrolled approximations in order to faithfully resolve ML CrI3’s remaining controversies and uncertainties.

FIG. 1.

Geometry of monolayer CrI3 cleaved from the bulk structure reported in Ref. 45. (a) Top view depicting a lattice constant of a0 = 6.867 Å and the bond angles θ1 and θ2 computed in this work. (b) Side view depicting the Cr1–I bond distance of 2.726 Å (purple) and the Cr1–Cr2 bond distance of 3.965 Å (blue).

FIG. 1.

Geometry of monolayer CrI3 cleaved from the bulk structure reported in Ref. 45. (a) Top view depicting a lattice constant of a0 = 6.867 Å and the bond angles θ1 and θ2 computed in this work. (b) Side view depicting the Cr1–I bond distance of 2.726 Å (purple) and the Cr1–Cr2 bond distance of 3.965 Å (blue).

Close modal

We utilize a combination of DFT with an on-site Hubbard-U correction (DFT + U) as well as many-body DMC simulations to resolve the structural, magnetic, and phonon properties of monolayer CrI3. Our results show that previous DFT predictions of ML CrI3’s magnetic spin moments48 are correct on average across a unit cell but miss critical local spatial fluctuations and anisotropies in the spin density. Our DMC calculations predict a magnetic moment of 3.62 μB per chromium atom and −0.145 μB per iodine atom, which is considerably larger than past estimates in the literature. We furthermore exploit a recently developed surrogate Hessian DMC line search technique42 to determine CrI3’s monolayer geometry with DMC accuracy, yielding high-accuracy bond lengths and bond angles that resolve previous structural ambiguities. Finally, our calculations also reveal that ML CrI3 possesses a substantial spin–phonon coupling, 3.32 cm−1, in line with couplings recently observed in other magnetic 2D materials. These findings indicate a far more anisotropic spin density involving more substantial ligand magnetism than previously thought.

In Sec. II, we describe the DFT and quantum Monte Carlo methods used, and in Sec. III, we present results on the monolayer’s structure, spin density, magnetic moments, and phonon spectra. Finally, Sec. IV contains a summary and conclusions.

The ground state properties of the CrI3 monolayer, including its lattice constant and magnetic moments, were modeled using DFT + U, Variational Monte Carlo (VMC), and DMC. DFT simulations were first performed to obtain reference structural data, guide our surrogate Hessian line search, and generate DFT + U wave functions that were subsequently used as inputs into progressively more accurate VMC and DMC simulations. VMC and DMC calculations were then performed to obtain CrI3’s geometry, spin density, and magnetic moments. In addition, DFT + U simulations with a self-consistent Hubbard-U correction53 were employed to investigate the effects of long-range magnetic ordering on lattice phonons where the range of U values considered was guided by our DMC findings.

DFT calculations were performed using the Vienna Ab Initio Simulation Package (VASP).55–57 The ion–electron interaction was described with the projector augmented wave (PAW) method.58,59 A cutoff energy of 500 eV was used for the plane-wave basis set. The Brillouin zone was sampled with an 8 × 8 × 1 Monkhorst–Pack k-point mesh for the unit cell of CrI3, which includes two Cr and six I atoms. A wide range of exchange correlation (XC) functionals were used to investigate the electronic and magnetic properties of the CrI3 monolayer. The Local Density Approximation (LDA), Perdew–Burke–Ernzerhof (PBE), PBE + U, and PBEsol60 functionals were specifically employed in this study. All calculations were spin-polarized. In its ferromagnetic (FM) configuration, all of CrI3’s magnetic moments were initialized in the same out-of-plane direction, while in its antiferromagnetic (AFM) configuration, the two Cr atoms per unit cell were set to have antiparallel spins. A large vacuum of more than 25 Å along the z-direction was employed to avoid artificial interactions between images. The energies were converged with a 1 × 10−8 eV tolerance. In addition to screening various XC functionals, a self-consistent Hubbard U (ULR) was determined from first principles by using the linear response approach proposed by Cococcioni and De Gironcoli,53 in which U was determined by the difference between the screened and the bare second derivative of the energy with respect to localized state occupations. The linear-response calculation was also performed using Quantum ESPRESSO, using the same high-quality (but hard) pseudopotentials as were used to generate trial wave functions for the Quantum Monte Carlo simulations (discussed below). We obtained a linear-response U (ULR) value of ∼3.3 eV for Cr with the LDA functional.

VMC and DMC calculations were undertaken using QMCPACK,44 a process that was facilitated by a Nexus workflow.61 All QMC calculations were performed using a norm-conserving scalar-relativistic Opium-generated Cr pseudopotential and a norm-conserving non-relativistic iodine pseudopotential that was developed by Burkatzki, Fillipi and Dolg.62 The comparison of DMC energies obtained using nonrelativistic and relativistic iodine pseudopotentials yielded an energy difference of 0.009 meV/f.u., which was deemed negligible for this study. Both pseudopotentials were of the Troullier–Martins flavor. Trial wave functions used as starting points for the VMC calculations were produced using LDA + U with U = 2 eV within Quantum ESPRESSO. In the DFT calculations, a plane-wave energy cutoff of 300 Ry was necessary to converge the total energy due to the high-quality (but hard) pseudopotentials. The trial Jastrow factor consisted of inhomogeneous one- and homogeneous two-body terms, each represented by sums of 1D B-spline-based correlation functions in electron–ion or electron–electron pair distances. The trial Jastrow factor was optimized using the linear method. The VMC energy and variance were converged at a B-spline meshfactor of 1.2 using the fixed optimized Jastrow factor.

To determine a time step that is a reasonable compromise between speed and accuracy, as well as the U in LDA + U trial wave functions that minimizes fixed-node errors, preliminary calculations were performed on a 2 × 2 × 1 supercell. These calculations demonstrated that a time step of 0.01 Ha−1 is capable of converging the DMC energies to within 0.01 Ha per formula unit (see Fig. 2). Moreover, as depicted in Fig. 3, the lowest DMC energies were achieved using a LDA + U trial wave function with U = 2 eV. As shown in Fig. SI, U = 2 eV also minimized the DMC energy when a smaller time step of 0.001 Ha−1 was employed, demonstrating that this U is robust to time step errors when sufficiently small time steps are used. A time step of 0.01 Ha−1 and LDA + U trial wave functions with U = 2 eV were thus used in all of our subsequent DMC production runs. Note that the obtained optimal value of U from DMC is close to that of the ULR ∼ 3.3 eV value obtained from linear-response calculations. Furthermore, both values agree with statistically indistinguishable total energies from a recent DMC study of bulk CrI3 (see Fig. 1 in Ref. 63). All of our calculations also utilized a 3 × 3 × 1 QMC twist grid and a meshfactor of 1.2.

FIG. 2.

Time step extrapolation of DMC energies obtained for a 2 × 2 × 1-tiled primitive cell of monolayer CrI3. The blue points represent DMC energies and their accompanying error bars computed for the given time step sizes (dt), while the red point and its error bar denote the extrapolated zero-time step energy based upon a linear fit of the data. The black line denotes the best linear fit to the data.

FIG. 2.

Time step extrapolation of DMC energies obtained for a 2 × 2 × 1-tiled primitive cell of monolayer CrI3. The blue points represent DMC energies and their accompanying error bars computed for the given time step sizes (dt), while the red point and its error bar denote the extrapolated zero-time step energy based upon a linear fit of the data. The black line denotes the best linear fit to the data.

Close modal
FIG. 3.

Twist-averaged DMC energies for a 2 × 2 × 1 supercell of monolayer CrI3 obtained using LDA + U trial wave functions with varying values of U. The minimum energy occurs around U = 2 eV, suggesting that a LDA + U trial wave function with U = 2 would be the best to use in production-level DMC calculations.

FIG. 3.

Twist-averaged DMC energies for a 2 × 2 × 1 supercell of monolayer CrI3 obtained using LDA + U trial wave functions with varying values of U. The minimum energy occurs around U = 2 eV, suggesting that a LDA + U trial wave function with U = 2 would be the best to use in production-level DMC calculations.

Close modal

VMC calculations employing twist-averaged boundary conditions were performed for increasingly larger twist grids and were shown to converge with a 3 × 3 × 1 twist grid; this size twist grid was thus used for twist-averaging of all subsequent VMC and DMC calculations to minimize one-body finite size effects.64 We further employed extrapolation to the thermodynamic limit of Model Periodic Coulomb-corrected and uncorrected DMC energies for increasingly larger supercells with 1 × 2 × 1 (16 atoms), 2 × 2 × 1 (32 atoms), 2 × 3 × 1 (48 atoms), and 3 × 3 × 1 (72 atoms) tilings to minimize the two-body finite-size errors (Fig. 4).44,65 All DMC runs used the T-moves scheme for pseudopotential evaluation to minimize localization errors.66,67

FIG. 4.

Finite size extrapolation of DMC energies per formula unit (f.u.) to their thermodynamic limit using 32 (2 × 2 × 1)-, 48 (2 × 3 × 1)-, and 72 (3 × 3 × 1)- atom supercells. The green circles denote energies corrected using model periodic Coulomb (MPC) finite size corrections, while the black circles denote uncorrected energies. Note that the finite size-extrapolated energies obtained with and without corrections are within 2 mHa of one another.

FIG. 4.

Finite size extrapolation of DMC energies per formula unit (f.u.) to their thermodynamic limit using 32 (2 × 2 × 1)-, 48 (2 × 3 × 1)-, and 72 (3 × 3 × 1)- atom supercells. The green circles denote energies corrected using model periodic Coulomb (MPC) finite size corrections, while the black circles denote uncorrected energies. Note that the finite size-extrapolated energies obtained with and without corrections are within 2 mHa of one another.

Close modal

To predict the site-averaged atomic magnetic moment per chromium, MCr, and iodine, MI, the spin densities, ρs(r), obtained from both our DFT and DMC simulations were integrated up to a cutoff radius rcut, defined as the zero-recrossing radius of the sign of the spin density (see Fig. SVI). In particular, we sum over the spherically interpolated spin densities within a 24-atom unit cell to predict the magnetic moment per atom (MA),

(1)

where the ri denote the distances from the center of the atom to the given point on the grid. To facilitate a direct comparison between the DFT and DMC spin densities, the spin densities obtained from DFT simulations were interpolated onto the dimensions of the DMC spin density grid, before being mapped onto a spherical grid.

In order to predict highly accurate lattice parameters for ML CrI3, we used a surrogate Hessian-accelerated DMC energy-based structural optimization method which has recently demonstrated robust performance when applied to two-dimensional GeSe.42 This method leverages the DFT energy Hessian, which is substantially cheaper to compute than the DMC energy Hessian, to direct where to compute DMC energies based on optimal statistical sampling to ultimately determine the lowest-energy material geometry. This line search method resolves structural parameters to an accuracy higher than that obtainable via DFT energy gradient-based structural optimization techniques; this resolution is particularly important for materials which exhibit sensitive coupling between electronic, magnetic, and structural degrees of freedom, such as 2D CrI3.

Our line search begins by first constructing an approximate energy Hessian (Hp), or a force constant matrix, using a cheaper theory, which in our case is DFT. This step is meant to (1) address the expense that would result from exploring an arbitrarily complex PES from scratch using only QMC energies and (2) minimize the noise that would result from numerous stochastic energy evaluations by allowing for a few optimally placed QMC energy calculations. We expanded the DFT PES to the second order in the Wyckoff parameter space, p, as in Ref. 68,

(2)

and diagonalized the Hessian to obtain search directions conjugate to the PES isosurfaces,

(3)

where the columns of UT form an optimal basis of parameter directions for the line search.

With these search directions, a set of structures containing an equilibrium structure and a total of 18 “strained” structures are generated, which comprise the structure population for the first line search iteration. More specifically, three increasingly “positively” (tensile) strained structures and three increasingly “negatively” (compressive) strained structures are generated along each search direction by systematically varying the corresponding parameters of the equilibrium structure. Next, DMC energies are obtained for the entire structure population. The shape of the resultant DMC PES is then used to locate a candidate minimum energy structure. The next iteration of the line search is performed identically to the first but starting from this new candidate minimum structure. This process is repeated until the statistical noise on the parameter error bars is within the desired limit.68 Like all conjugate direction methods, a property of the method is the potential to converge after only a single iteration within a suitably quadratic region of the PES and using ideal search directions. In this work, converged DMC-based structural parameters were obtained after only three iterations and were averaged over the last two iterations to account for remnant statistical fluctuations in the lattice constant.

Phonon calculations meant to examine monolayer CrI3’s spin–phonon coupling were performed using the frozen phonon method as implemented in the PHONOPY code69 based upon DFT calculations performed in VASP.57,59 We explored how CrI3’s phonons change with different magnetic orderings, including nonmagnetic, ferromagnetic, and antiferromagnetic orderings, using both the LDA and LDA + U functionals. Our choice of U = 3 eV is further motivated below. 2 × 2 × 1 supercell structures with displacements were created from a unit cell fully preserving the material’s crystal symmetry. To properly analyze how the phonons vary in different magnetic phases, the supercell studied must be commensurate with the magnetic ordering. The 2 × 2 × 1 supercell selected was large enough to be commensurate with the FM and in-plane AFM ordering considered in this paper. Force constants were calculated using the structure files from the computed forces on the atoms. A part of the dynamical matrix was built from the force constants. Phonon frequencies and eigenvectors were calculated from the dynamical matrices with the specified q-points. Monolayer CrI3 possesses D3d point group symmetry, and hence, the phonon modes at the Γ point can be decomposed as GD3d = 2A1g + 2A2g + 4Eg + 2A1u + 2A2u + 4Eu. Excluding the three acoustic modes (the doubly-degenerate Eu and A1u modes) and noticing that each of the Eg and Eu modes is doubly degenerate, there are 21 modes in total.

As a starting point for our subsequent DMC calculations, we began by modeling the monolayer’s properties using standard density functionals, including the LDA, the GGA implementation of Perdew–Burke–Ernzenhof (PBE),72 PBEsol,73 and a series of PBE + U functionals. Previous work has shown that the material properties can depend strongly on the functional employed, and thus, we initially sought a functional that could simultaneously predict the lattice constant, magnetic moment, and electronic bandgap of the monolayer. Since relatively few ab initio and experimental studies have been able to ascertain the properties of the monolayer, we fixed the internal geometry of the monolayer to the previously determined bulk values. As illustrated in Fig. 5, no single functional came close to simultaneously reproducing all three bulk properties: PBE + U = 1 eV most closely approximated the bulk lattice constant, PBE + U = 6 eV most closely approximated the bulk magnetic moment, and PBEsol most closely approximated the bulk bandgap. Even more glaringly, all three properties continuously increased with the U employed in our PBE + U calculations, strongly suggesting that the U’s employed could be continuously tuned to reproduce any individual quantity desired. This is illustrative of the sensitivity of CrI3 properties to the variability in DFT functionals, which motivates the use of DMC throughout the rest of this work.

FIG. 5.

DFT predictions of the lattice constant, magnetic moment, and electronic bandgap of bulk CrI3 using different functionals. In the above, Ef.u. designates the energy per formula unit.

FIG. 5.

DFT predictions of the lattice constant, magnetic moment, and electronic bandgap of bulk CrI3 using different functionals. In the above, Ef.u. designates the energy per formula unit.

Close modal

To select a DFT functional that could serve as a meaningful reference, we thus turned to DMC simulations. Based on Fig. 3, LDA + U wave functions with U’s of 2–3 eV minimize the DMC energy, so in all subsequent DFT analyses, we have employed LDA + U wave functions with U = 2–3 eV. This range of U values is consistent with our linear-response estimate of ULR = 3.3 eV and also that employed in a wide range of materials and a recent DMC study of bulk CrI3.63 

With a U = 2 eV value, we used LDA + U to relax the CrI3 structure to determine its equilibrium geometry. As shown in Tables I and SI, LDA + U yields a lattice constant of 6.695 Å and an axial iodine–chromium–iodine bond angle (θICrI) of 175.72°. The predicted lattice constant is within the range of lattice constants previously obtained using other density functionals, which range from 6.686 Å from previous DFT + U calculations46 to 6.978 Å from GGA + U calculations71 (a variation of roughly 5%). In contrast, recent STM experiments point to an experimental monolayer lattice parameter of 6.84 Å, which is greater than our best LDA + U estimate (a point to which we will return below).

Given this variability in DFT-derived lattice constants, we employed DMC guided by a surrogate Hessian method to determine monolayer CrI3’s equilibrium geometry with DMC-level accuracy. For monolayer CrI3, the surrogate Hessian method yielded converged structural parameters to within the desired accuracy of 0.5% of the lattice parameter within three iterations.

As evidenced by the tightly bunched contour lines in Fig. 6, the potential energy surface around the DMC structural minimum is steep and highly harmonic. In terms of the Wyckoff parameters x and a (note that aa0), the DMC PES is more shallow along the direction in which x and a simultaneously increase and steeper in the direction in which x increases but a decreases. In contrast, the PES is steeper along the direction in which z and a simultaneously increase and more shallow along the direction in which z decreases but a increases. Finally, although the relationship between z and x is slightly more anharmonic than the previous two relationships, the PES roughly tends to be more shallow as both z and x increase, but steeper in the direction in which z decreases, while x increases.

FIG. 6.

DMC potential energy surfaces obtained using the surrogate Hessian approach. The PES along the (a) (a, x) parameter pair plane, (b) (a, z) parameter pair plane, and (c) (x, z) parameter pair plane. Here, a, x, and z denote Wyckoff parameters, of which aa0 in Bohr and x and z combine to give bond angles and distances. The energy contours are separated by 4 mHa.

FIG. 6.

DMC potential energy surfaces obtained using the surrogate Hessian approach. The PES along the (a) (a, x) parameter pair plane, (b) (a, z) parameter pair plane, and (c) (x, z) parameter pair plane. Here, a, x, and z denote Wyckoff parameters, of which aa0 in Bohr and x and z combine to give bond angles and distances. The energy contours are separated by 4 mHa.

Close modal

As bond angles and lengths consist of contributions from both the x and z Wyckoff parameters, a physical understanding of the DMC PES is enhanced by considering Fig. 7 of the main text and Figs. SII and SIII of the supplementary material. In these figures, slices of the DMC PES near the true minima are taken so as to illustrate (i) how the CrI3 energy changes with the lattice constant, a0; Cr–I bond length, dCrI; and CrI3 bond angles, θ1 and θ2 (see Fig. 1 for a visualization of how these quantities are defined) and (ii) that these structural quantities differ significantly between LDA + U and DMC. These figures reveal that the energy changes most rapidly as the lattice constant is varied. The energy minima become increasingly more shallow along the dCrI, θ1, and θ2 directions, signifying that it is more energetically costly to isotropically strain the lattice than to more locally vary bond lengths and angles. The relatively low barriers to changing the bond distances and angles are what, moreover, make it challenging to accurately resolve CrI3’s structure, a feature that CrI3 holds in common with many 2D materials, as we have illustrated using DMC in our previous works.42 

FIG. 7.

Relative energies for ML CrI3 obtained using DMC and LDA + U (U = 2 eV) vs (a) the lattice constant and (b) the Cr–I bond distance. E0 denotes the minimum energy per formula unit (f.u.) yielded by each method, with LDA + U E0 = −185.3 Ha and DMC E0 = −121.1 Ha. The optimal lattice constant and Cr–I bond distance obtained by DMC are denoted by vertical orange solid lines, while those obtained by LDA + U (U = 2 eV) are denoted by orange dashed lines.

FIG. 7.

Relative energies for ML CrI3 obtained using DMC and LDA + U (U = 2 eV) vs (a) the lattice constant and (b) the Cr–I bond distance. E0 denotes the minimum energy per formula unit (f.u.) yielded by each method, with LDA + U E0 = −185.3 Ha and DMC E0 = −121.1 Ha. The optimal lattice constant and Cr–I bond distance obtained by DMC are denoted by vertical orange solid lines, while those obtained by LDA + U (U = 2 eV) are denoted by orange dashed lines.

Close modal

Ultimately, our line search converges upon an LS-DMC lattice parameter of a0 = 6.87 Å, as presented in Fig. 7 and Table I. This lies within the range of previous DFT-derived estimates and is 2% larger than the DFT + U lattice constant discussed earlier. Nevertheless, what adds particular confidence to these results is that this independently derived lattice parameter is just 0.4% off from that recently obtained for a CrI3 monolayer grown using molecular beam epitaxy and analyzed using scanning tunneling microscopy (see Li et al. in Table I).32 The fact that our surrogate Hessian approach was able to so closely reproduce an experimental value underscores both DMC and the surrogate Hessian approach’s accuracy for CrI3. In addition to obtaining the monolayer lattice parameter, these calculations also yielded estimates of the chromium–iodine bond distance (dCrI = 2.723 Å), as also presented in Fig. 7, and the monolayer bond angles (θ1 = 90.4° and θ2 = 175.4°), as tabulated in Table SI and Figs. SII and SIII of the supplementary material.

TABLE I.

Tabulation of lattice parameters and magnetic moments predicted in this work compared to those from other studies (* experimentally determined in study of ML CrI3 in Ref. 32).

ReferencesMethoda0 (Å)mCr (μB)mI (μB)
This work LS-DMC 6.87(3) 3.61(9) −0.14(5) 
This work LDA + U 6.695 3.497 −0.099 
Li et al.32  GGA + U 6.84* 3.28 ⋯ 
Yang et al.70  GGA + U ⋯ 3.32 ⋯ 
Wu et al.71  GGA + U 6.978 3.106 ⋯ 
Lado and DFT + U 6.686 ⋯ 
Fernández-Rossier46     
Zhang et al.51  PBE(HSE06) 7.008 3.103 ⋯ 
ReferencesMethoda0 (Å)mCr (μB)mI (μB)
This work LS-DMC 6.87(3) 3.61(9) −0.14(5) 
This work LDA + U 6.695 3.497 −0.099 
Li et al.32  GGA + U 6.84* 3.28 ⋯ 
Yang et al.70  GGA + U ⋯ 3.32 ⋯ 
Wu et al.71  GGA + U 6.978 3.106 ⋯ 
Lado and DFT + U 6.686 ⋯ 
Fernández-Rossier46     
Zhang et al.51  PBE(HSE06) 7.008 3.103 ⋯ 

Although bond angles and distances have not been measured experimentally in the monolayer and are sparsely reported in DFT studies, our line search-predicted values all fall within 0.5% of experimental values for bulk CrI3. The other bond angles and distances obtained using different DFT functionals tabulated in Table SI of the supplementary material differ from the experimental bulk values by up to 3%, further highlighting the robustness of our surrogate Hessian line search method for predicting structural parameters.

Using our DMC-optimized CrI3 structure, we then proceeded to compute the magnetic moments on the Cr and I atoms in the hopes of fully and accurately resolving the magnetic structure of the monolayer for the first time. By integrating over the DMC spin density, as described in Sec. II C, we obtained a magnetic moment of 3.62 μB on each Cr and a moment of −0.15 μB on each I. Interestingly, this indicates that each Cr possesses a magnetic moment significantly larger than that of the +3 charge that one would naively assume based upon chromium’s typical oxidation state. Many previous DFT studies obtained an average magnetic moment of 3 μB across the entire CrI3 unit cell, leading many researchers to conclude that all of the unit cell’s magnetism could be attributed to that from the Cr site.

While the sum of our magnetic moments for the Cr and three I atoms that constitute the unit cell also totals to 3 μB, we instead find that the iodine atoms additionally carry a significant magnetic moment, suggesting that CrI3’s magnetism is more complex—and nonlocal—than previously assumed. As a check on these moments, we also employed LDA + U and DMC to compute the moments of a monolayer cleaved from the experimental bulk structure.45 The moments obtained using this structure were different by up to 0.06 μB, supporting the assertion that DMC is capturing correlations missed by DFT (Fig. 8). Furthermore, the DMC moments obtained using this structure were within 0.005 μB of those from our line search-optimized structure (see Table I), a statistically insignificant discrepancy that strongly corroborates our findings.

FIG. 8.

16-atom ML CrI3 spin densities [isosurface = 0.000 15 (nupndown)/a03] for a monolayer cleaved from an experimental structure45 obtained with (a) LDA + U (U = 2 eV) and (b) DMC. Magnetic moments of Cr as a function of the distance, rcut, from the center of the Cr atom, obtained by interpolating over the (c) LDA + U (U = 2 eV) and (d) DMC spin density grids. Areas with a net positive spin density are colored yellow, while those with a net negative spin density are colored turquoise. The medium blue coloring is an artifact of the viewing angle. Magnetic moments are depicted by the orange lines, which are positioned at the maxima of the magnetic moment curves.

FIG. 8.

16-atom ML CrI3 spin densities [isosurface = 0.000 15 (nupndown)/a03] for a monolayer cleaved from an experimental structure45 obtained with (a) LDA + U (U = 2 eV) and (b) DMC. Magnetic moments of Cr as a function of the distance, rcut, from the center of the Cr atom, obtained by interpolating over the (c) LDA + U (U = 2 eV) and (d) DMC spin density grids. Areas with a net positive spin density are colored yellow, while those with a net negative spin density are colored turquoise. The medium blue coloring is an artifact of the viewing angle. Magnetic moments are depicted by the orange lines, which are positioned at the maxima of the magnetic moment curves.

Close modal

In contrast, integration over the spin densities of the LDA + U-optimized structure yielded magnetic moments of 3.497 μB on each Cr and −0.099 μB on each I, meaning that the correlation accounted for by DMC calculations leads to an increased spin anisotropy in the unit cell. That correlation leads to an increase in spin anisotropy is further substantiated by the fact that we see an increase in the magnitudes of the moments in the DMC-optimized structure to 3.655 μB on each Cr and −0.153 μB on each I after applying VMC to our DFT trial wave function, followed by a further tuning to their final values upon applying DMC to our VMC trial wave function (see Fig. SIV of the supplementary material).

Finally, we used LDA + U = 3 eV in VASP to calculate atomic magnetic moments with and without spin–orbit coupling (SOC) to gauge the effect of SOC on the atomic magnetic moments. When SOC was turned off, we obtained moments of 2.848 μB and −0.073 μB on the Cr and I, respectively. When the SOC was turned on, the magnetic moment on the Cr grew to 2.856 μB, while the magnetic moment on the I remained −0.073 μB. This suggests that leaving SOC out of our DMC calculations should not drastically affect our magnetic moment results.

The discrepancies in the magnetic moments that we observe with varying monolayer geometries point to appreciable spin–phonon and spin–lattice couplings. Spin–phonon and spin–lattice couplings are associated with changes in the magnetic exchange interaction with changes in ionic motion or strain, respectively, and together can give rise to a large magneto-elastic effect. A material with a large spin–phonon coupling can be leveraged for magneto-caloric device applications.

Up to the lowest order, the shift in the phonon frequency due to coupling to the lattice can be attributed to ν = ν0 + λSi · Sj⟩, where ν0 is the frequency in the paramagnetic case and λ is the spin–phonon coupling constant with i and j as site indices. For CrI3, if we assume a nominal +3 oxidation state for Cr, the expectation value of this spin–spin correlation function equates to 9/4. Hence, for a given phonon frequency shift, we can estimate λ. The spin–spin correlation function is a constant for all modes in a given material. Hence, we simply report the frequency shifts in much of our discussion below. Nevertheless, we would like to note that for a given frequency shift, the effects of spin-anisotropy due to strong correlation, as revealed by our DMC calculations, should significantly increase the spin–spin correlation function, thereby effectively reducing the spin–phonon coupling.

Experimentally, the largest spin–phonon coupling is observed in a 5d perovskite oxide, corresponding to a frequency shift of 40 cm−1.74 In comparison, a frequency shift of ∼2.7 cm−1 is observed for the Eg mode of the chromium-based 2D material Cr2Ge2Te6 (CGT) around its Tc. This corresponds to a λ of ∼1.2 cm−1. Spin–phonon couplings in the range of 0.27–0.4 cm−1 have subsequently been observed in other transition-metal trihalides, again arising from the Eg modes.75 

In our phonon calculations, we initially fixed CrI3’s lattice parameter to its bulk value. Considering AFM ordering as one specific realization of the paramagnetic phase, we find that the largest change in phonon frequency between the AFM and FM phases is for the Eg symmetric mode and is ∼4 cm−1 (see Table II, modes 8 and 9). This is consistent with CrI3 having a sizable spin–phonon coupling at finite temperatures, comparable to that of CGT and other 2D layered magnetic materials.

TABLE II.

Frequencies of the phonon modes for the FM, AFM, and NM phases of the CrI3 monolayer using the bulk and relaxed (r) lattice parameters predicted with the LDA + U = 3 eV functional in VASP. Frequencies are specified in units of cm−1.

ModeSymm.FMFMrAFMAFMrNMNMr
1, 2 Eg 47.6 50.8 47.3 49.6 39.9, 40.4 19.1, 41.6 
A2u 49.7 57.1 50.4 58.8 52.2 47.7 
A1g 69.6 76.9 69.8 77.8 53.2 53.6 
5, 6 Eu 78.2 81.2 79.2 82.9 74.5, 87.6 60.8, 71.8 
A2g 84.5 88.7 86.3 90.7 91.4 82.6 
8, 9 Eg 102 103 97.9 92.5 94.7, 104 92.4, 94.7 
10, 11 Eg 107 109 108 110 107, 110 103, 112 
12, 13 Eu 109 116 111 118 115, 120 115, 118 
14 A1g 129 131 126 128 140 173 
15 A2u 130 135 131 136 162 194 
16 A2g 212 219 218 222 228 205 
17, 18 Eu 221 229 227 235 231, 232 206, 222 
19, 20 Eg 238 245 236 235 241, 251 236, 254 
21 A1u 258 267 251 261 313 298 
ModeSymm.FMFMrAFMAFMrNMNMr
1, 2 Eg 47.6 50.8 47.3 49.6 39.9, 40.4 19.1, 41.6 
A2u 49.7 57.1 50.4 58.8 52.2 47.7 
A1g 69.6 76.9 69.8 77.8 53.2 53.6 
5, 6 Eu 78.2 81.2 79.2 82.9 74.5, 87.6 60.8, 71.8 
A2g 84.5 88.7 86.3 90.7 91.4 82.6 
8, 9 Eg 102 103 97.9 92.5 94.7, 104 92.4, 94.7 
10, 11 Eg 107 109 108 110 107, 110 103, 112 
12, 13 Eu 109 116 111 118 115, 120 115, 118 
14 A1g 129 131 126 128 140 173 
15 A2u 130 135 131 136 162 194 
16 A2g 212 219 218 222 228 205 
17, 18 Eu 221 229 227 235 231, 232 206, 222 
19, 20 Eg 238 245 236 235 241, 251 236, 254 
21 A1u 258 267 251 261 313 298 

We next performed full geometry relaxations (using LDA + U = 3 eV in VASP) for the different magnetic configurations. The DMC calculations depicted in Fig. 3 helped us to narrow the range of U’s to employ in our phonon calculations down to U = 2–3. We ultimately employed U = 3 to obtain the data in Table II. Phonon frequencies obtained using U = 2 differed by at most 2.5 cm−1 from those reported in the table, supporting this decision (see Table SII). We find the optimal lattice parameters for the FM, AFM, and NM phases of the CrI3 monolayer to be 6.677, 6.656, and 6.617 Å, respectively (see Table II). A large, ∼1%, difference in lattice-parameters due to magnetic ordering is indicative of a non-trivial spin–lattice coupling and is consistent with unstable in-plane acoustic modes when ML-CrI3 is forced to be in a non-magnetic state (see Fig. S5). Unstable acoustic modes have been observed in certain Heusler compounds.76 A strong spin–lattice coupling, particularly involving soft in-plane acoustic modes, should lead to structural transitions across the Curie temperature. Indeed, bulk CrI3 has been experimentally shown to transition from a rhombohedral to a monoclinic structure across the paramagnetic transition,45 consistent with these findings.

We also find that the phonon frequencies shift to larger values across all of the different magnetic orderings when the lattice parameter is allowed to relax, as shown in Table II. Specifically, the phonon frequency change between the FM and AFM phases increases to ∼10 cm−1 for the Eg mode (modes 8 and 9). This particular mode resembles shearing of the iodine planes and is the one with the largest spin–phonon coupling in CGT as well as in other Cr trihalides. Assuming a spin-eigenvalue S of 3.47/2 based upon our DMC calculations of the magnetic moment, this corresponds to a λ of ∼3.32 cm−1. In comparison, the theoretically predicted spin–phonon coupling for CGT using a rigorous perturbation theory approach was 3.19 cm−1.77 

These results demonstrate that even in the monolayer limit, CrI3 possesses strong spin–phonon and spin–lattice couplings. While recent studies have demonstrated a magneto-optic effect in few-layer CrI3, the large coupling of the magnetism to the Raman active phonon frequencies and lattice-strain that we observe in this study suggest that one could potentially use magnetic fields to modulate inelastically scattered light. Indeed, a very recent experiment demonstrates that an out-of-plane magnetic field change from −2.5 to 2.5 T leads to a rotation in the plane of the polarization of inelastically scattered light from −20° to +60°.78 

In summary, we have used a potent combination of first principles DFT and DMC calculations to produce some of the most accurate estimates of the electronic, magnetic, and structural properties of monolayer CrI3 to date. Using a surrogate Hessian line search optimization technique combined with DMC, we were able to independently resolve the lattice and other structural parameters of CrI3 to within a fraction of a percent of recently published STM measurements, an accomplishment given the up to 10% variability in previous DFT-derived estimates of the lattice parameter, depending upon the functional employed. Based upon the DMC-quality structure we obtained, we were then able to acquire a high-resolution monolayer spin density that showed each Cr atom to possess a magnetic moment of 3.62 μB and each I atom to have a moment of −0.145 μB, substantially larger moments than previously reported values that suggest that CrI3 manifests substantial ligand magnetism. Given the Cr–I–Cr angle of 90°, this would indicate a superexchange stabilization of the ferromagnetic ordering. In conjunction with the expected large spin–orbit coupling on the I atom,46 this could also give rise to a large ligand superexchange-dominated magnetic-anisotropy,79,80 explaining recent observations of spin-waves and magnons in thin-film CrI3.81 We, moreover, demonstrated that CrI3 possesses remarkably strong spin–phonon coupling, with a predicted value as large as 3.32 cm−1. This work, thus, demonstrates CrI3’s promise for magnon-based spintronic applications,82 potentially with optical controls,78 while also demonstrating the capability of DMC to accurately model the structural and magnetic properties of 2D materials.

See the supplementary material for additional figures and tables describing the time step extrapolations performed, lattice parameters, and magnetic moment calculations.

The authors thank Paul Kent, Kemp Plumb, Anand Bhattacharya, Ho Nyung Lee, Fernando Reboredo, Nikhil Sivadas, and the QMCPACK team for thoughtful conversations. The work by G.H., J.T., R.N., J.K., M.C.B., O.H., P.G., and B.R.; modeling and analysis efforts by D.S.; and the scientific applications of QMCPACK were supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, as part of the Computational Materials Sciences Program and the Center for Predictive Simulation of Functional Materials. The preparation of this manuscript by D.S. was supported by the NASA Rhode Island Space Grant Consortium. This research was conducted using computational resources and services at the Center for Computation and Visualization, Brown University, and the National Energy Research Scientific Computing Center (NERSC), a U.S. Department of Energy Office of Science User Facility operated under Contract No. DE-AC02-05CH11231.

The authors have no conflicts to disclose.

D.S. and G.H. made equal contributions to the paper.

The data that support the findings of this study are openly available in Blaiszik https://doi.org/10.1007/s11837-016-2001-3, Ref. 83, Blaiszik https://doi.org/10.1557/mrc.2019.118, Ref. 84 and the supplementary materials.

1.
K. S.
Novoselov
,
A.
Mishchenko
,
A.
Carvalho
, and
A. H.
Castro Neto
, “
2D materials and van der Waals heterostructures
,”
Science
353
,
aac9439
(
2016
).
2.
K.
Tang
and
W.
Qi
, “
Moiré-pattern-tuned electronic structures of van der Waals heterostructures
,”
Adv. Funct. Mater.
30
,
2002672
(
2020
).
3.
F.
He
,
Y.
Zhou
,
Z.
Ye
,
C.
Sang-Hyeok
,
J.
Jeong
,
X.
Meng
, and
Y.
Wang
, “
Moiré patterns in 2D materials: A review
,”
ACS Nano
15
,
5944
(
2021
).
4.
D.
Qiu
,
C.
Gong
,
S.
Wang
,
M.
Zhang
,
C.
Yang
,
X.
Wang
, and
J.
Xiong
, “
Recent advances in 2D superconductors
,”
Adv. Mater.
33
,
2006124
(
2021
).
5.
X.
Chen
,
J. L.
Schmehr
,
Z.
Islam
,
Z.
Porter
,
E.
Zoghlin
,
K.
Finkelstein
,
J. P. C.
Ruff
, and
S. D.
Wilson
, “
Unidirectional spin density wave state in metallic (Sr1−xLax)2IrO4
,”
Nat. Commun.
9
,
103
(
2018
).
6.
H.
Huang
,
S. J.
Lee
,
Y.
Ikeda
,
T.
Taniguchi
,
M.
Takahama
,
C. C.
Kao
,
M.
Fujita
, and
J. S.
Lee
, “
Two-dimensional superconducting fluctuations associated with charge-density-wave stripes in La1.87Sr0.13Cu0.99Fe0.01O4
,”
Phys. Rev. Lett.
126
,
167001
(
2021
).
7.
L.
Wang
,
Y.
Wu
,
Y.
Yu
,
A.
Chen
,
H.
Li
,
W.
Ren
,
S.
Lu
,
S.
Ding
,
H.
Yang
,
Q. K.
Xue
,
F. S.
Li
, and
G.
Wang
, “
Direct observation of one-dimensional Peierls-type charge density wave in twin boundaries of monolayer MoTe2
,”
ACS Nano
14
,
8299
(
2020
).
8.
H. W.
Guo
,
Z.
Hu
,
Z. B.
Liu
, and
J. G.
Tian
, “
Stacking of 2D materials
,”
Adv. Funct. Mater.
31
,
2007810
(
2020
).
9.
N.
Sivadas
,
S.
Okamoto
,
X.
Xu
,
C. J.
Fennie
, and
D.
Xiao
, “
Stacking-dependent magnetism in bilayer CrI3
,”
Nano Lett.
18
,
7658
(
2018
).
10.
D.
Soriano
,
C.
Cardoso
, and
J.
Fernández-Rossier
, “
Interplay between interlayer exchange and stacking in CrI3 bilayers
,”
Solid State Commun.
299
,
113662
(
2019
).
11.
W.
Chen
,
X.
Gui
,
L.
Yang
,
H.
Zhu
, and
Z.
Tang
, “
Wrinkling of two-dimensional materials: Methods, properties and applications
,”
Nanoscale Horiz.
4
,
291
(
2018
).
12.
Z.
Dai
,
L.
Liu
, and
Z.
Zhang
, “
Strain engineering of 2D materials: Issues and opportunities at the interface
,”
Adv. Mater.
31
,
1805417
(
2019
).
13.
T.
Song
,
Z.
Fei
,
M.
Yankowitz
,
Z.
Lin
,
Q.
Jiang
,
K.
Hwangbo
,
Q.
Zhang
,
B.
Sun
,
T.
Taniguchi
,
K.
Watanabe
,
M. A.
McGuire
,
D.
Graf
,
T.
Cao
,
J.-H.
Chu
,
D. H.
Cobden
,
C. R.
Dean
,
D.
Xiao
, and
X.
Xu
, “
Switching 2D magnetic states via pressure tuning of layer stacking
,”
Nat. Mater.
18
,
1298
(
2019
).
14.
T.
Li
,
S.
Jiang
,
N.
Sivadas
,
Z.
Wang
,
Y.
Xu
,
D.
Weber
,
J. E.
Goldberger
,
K.
Watanabe
,
T.
Taniguchi
,
C. J.
Fennie
,
K.
Fai Mak
, and
J.
Shan
, “
Pressure-controlled interlayer magnetism in atomically thin CrI3
,”
Nat. Mater.
18
,
1303
(
2019
).
15.
G.
Trambly de Laissardière
,
D.
Mayou
, and
L.
Magaud
, “
Localization of Dirac electrons in rotated graphene bilayers
,”
Nano Lett.
10
,
804
808
(
2010
).
16.
R.
Bistritzer
and
A. H.
MacDonald
, “
Moiré bands in twisted double-layer graphene
,”
Proc. Natl. Acad. Sci. U. S. A.
108
,
12233
12237
(
2011
).
17.
J. M. B.
Lopes dos Santos
,
N. M. R.
Peres
, and
A. H.
Castro Neto
, “
Continuum model of the twisted graphene bilayer
,”
Phys. Rev. B
86
,
155449
(
2012
).
18.
M.
Yankowitz
,
S.
Chen
,
H.
Polshyn
,
Y.
Zhang
,
K.
Watanabe
,
T.
Taniguchi
,
D.
Graf
,
A. F.
Young
, and
C. R.
Dean
, “
Tuning superconductivity in twisted bilayer graphene
,”
Science
363
,
1059
1064
(
2019
).
19.
M.
Zeng
,
Y.
Xiao
,
J.
Liu
,
K.
Yang
, and
L.
Fu
, “
Exploring two-dimensional materials toward the next-generation circuits: From monomer design to assembly control
,”
Chem. Rev.
118
,
6236
(
2018
).
20.
M.
Gibertini
,
M.
Koperski
,
A. F.
Morpurgo
, and
K. S.
Novoselov
, “
Magnetic 2D materials and heterostructures
,”
Nat. Nanotechnol.
14
,
408
(
2019
).
21.
N. D.
Mermin
and
H.
Wagner
, “
Absence of ferromagnetism or antiferromagnetism in the one- or two-dimensional isotropic Heisenberg models
,”
Nat. Lett.
17
,
1307
(
1966
).
22.
P. C.
Hohenberg
, “
Existence of long-range order in one and two dimensions
,”
Phys. Rev.
158
,
383
(
1967
).
23.
B.
Huang
,
G.
Clark
,
E.
Navarro-Moratalla
,
D. R.
Klein
,
R.
Cheng
,
K. L.
Seyler
,
D.
Zhong
,
E.
Schmidgall
,
M. A.
McGuire
,
D. H.
Cobden
,
W.
Yao
,
D.
Xiao
,
P.
Jarillo-Herrero
, and
X.
Xu
, “
Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit
,”
Nat. Lett.
546
,
270
(
2017
).
24.
Z.
Fei
,
B.
Huang
,
P.
Malinowski
,
W.
Wang
,
T.
Song
,
J.
Sanchez
,
W.
Yao
,
D.
Xiao
,
X.
Zhu
,
A. F.
May
 et al, “
Two-dimensional itinerant ferromagnetism in atomically thin Fe3 GeTe2
,”
Nat. Mater.
17
,
778
782
(
2018
).
25.
X.
Wang
,
D.
Li
,
Z.
Li
,
C.
Wu
,
C.-M.
Che
,
G.
Chen
, and
X.
Cui
, “
Ferromagnetism in 2D vanadium diselenide
,”
ACS Nano
15
,
16236
(
2021
).
26.
A. K.
Behera
,
S.
Chowdbury
, and
S. R.
Das
, “
Magnetic skyrmions in atomic thin CrI3 monolayer
,”
Appl. Phys. Lett.
114
,
232402
(
2019
).
27.
J.
Liu
,
M.
Shi
,
P.
Mo
, and
J.
Lu
, “
Electrical-field-induced magnetic Skyrmion ground state in a two-dimensional chromium tri-iodide ferromagnetic monolayer
,”
AIP Adv.
8
,
055316
(
2018
).
28.
B.
Karpiak
,
A. W.
Cummings
,
K.
Zollner
,
M.
Vila
,
D.
Khokgriakov
,
A. M.
Hoque
,
A.
Dankert
,
P.
Svedlindh
,
J.
Fabian
,
S.
Roche
, and
S. P.
Dash
, “
Magnetic proximity in a van der Waals heterostructure of magnetic insulator and graphene
,”
2D Mater.
7
,
015026
(
2020
).
29.
M.
Onga
,
Y.
Sugita
,
T.
Ideue
,
Y.
Nakagawa
,
R.
Suzuki
,
Y.
Motome
, and
Y.
Iwasa
, “
Antiferromagnet-semiconductor van der Waals heterostructures: Interlayer interplay of exciton with magnetic ordering
,”
Nano Lett.
20
,
4625
(
2020
).
30.
A.
Mukherjee
,
K.
Shayan
,
L.
Li
,
J.
Shan
,
K. F.
Mak
, and
A. N.
Vamivakas
, “
Observation of site-controlled localized charge excitons in CrI3/WSe2 heterostructures
,”
Nat. Commun.
11
,
5502
(
2020
).
31.
W.
Jin
,
H. H.
Kim
,
Z.
Ye
,
G.
Ye
,
L.
Rojas
,
X.
Luo
,
B.
Yang
,
F.
Yin
,
J. S. A.
Horng
,
S.
Tian
,
Y.
Fu
,
G.
Xu
,
H.
Deng
,
H.
Lei
,
A. W.
Tsen
,
K.
Sun
,
R.
He
, and
L.
Zhao
, “
Observation of the polaronic character of excitons in a two-dimensional semiconducting magnet CrI3
,”
Nat. Commun.
11
,
4780
(
2020
).
32.
P.
Li
,
C.
Wang
,
J.
Zhang
,
S.
Chen
,
D.
Guo
,
W.
Ji
, and
D.
Zhong
, “
Single-layer CrI3 grown by molecular beam epitaxy
,”
Sci. Bull.
65
,
1064
(
2020
).
33.
G. H.
Ahn
,
M.
Amani
,
H.
Rasool
,
D. H.
Lien
,
J. P.
Mastandrea
,
J. W.
Ager
 III
,
M.
Dubey
,
D. C.
Chrzan
,
A. M.
Minor
, and
A.
Javey
, “
Strain-engineered growth of two-dimensional materials
,”
Nat. Commun.
8
,
608
(
2017
).
34.
Y.
Yan
,
H.
Liu
,
Y.
Han
,
F.
Li
, and
C.
Gao
, “
Substrate-affected lattice structural evolution in compressed monolayer ReS2
,”
Phys. Chem. Chem. Phys.
20
,
24927
(
2018
).
35.
M.
Velicky
and
P. S.
Toth
, “
From two-dimensional materials to their heterostructures: An electrochemist’s perspective
,”
Appl. Mater. Today
8
,
68
(
2017
).
36.
S.
Paul
,
S.
Haldar
,
S.
von Malottki
, and
S.
Heinze
, “
Role of higher-order exchange interactions for skyrmion stability
,”
Phys. Chem. Chem. Phys.
11
,
4756
(
2020
).
37.
I. V.
Soldatov
,
W.
Jiang
,
S. G. E.
Te Velthuis
,
A.
Hoffmann
, and
R.
Schäfer
, “
Size analysis of sub-resolution objects by Kerr microscopy
,”
Appl. Phys. Lett.
112
,
262404
(
2018
).
38.
L.
Thiel
,
Z.
Wang
,
M. A.
Tschudin
,
D.
Rohner
,
I.
Gutiérrez-Lezama
,
N.
Ubrig
,
M.
Gibertini
,
E.
Giannini
,
A. F.
Morpurgo
, and
P.
Maletinsky
, “
Probing magnetism in 2D materials at the nanoscale with single-spin microscopy
,”
Science
364
,
973
(
2019
).
39.
R.
Parr
and
Y.
Weitao
,
Density-Functional Theory of Atoms and Molecules, International Series of Monographs on Chemistry
(
Oxford University Press
,
1989
).
40.
D.
Wines
,
K.
Saritas
, and
C.
Ataca
, “
A first-principles quantum Monte Carlo study of two-dimensional (2D) GaSe
,”
J. Chem. Phys.
153
,
154704
(
2020
).
41.
E.
Mostaani
,
N. D.
Drummond
, and
V. I.
Fal’ko
, “
Quantum Monte Carlo calculation of the binding energy of bilayer graphene
,”
Phys. Rev. Lett.
115
,
115501
(
2015
).
42.
H.
Shin
,
J. T.
Krogel
,
K.
Gasperich
,
P. R. C.
Kent
,
A.
Benali
, and
O.
Heinonen
, “
Optimized structure and electronic band gap of monolayer GeSe from quantum Monte Carlo methods
,”
Phys. Rev. Mater.
5
,
024002
(
2021
).
43.
L.
Shulenburger
,
A. D.
Baczewski
,
Z.
Zhu
,
J.
Guan
, and
D.
Tománek
, “
The nature of the interlayer interaction in bulk and few-layer phosphorus
,”
Nano Lett.
15
,
8170
8175
(
2015
).
44.
J.
Kim
,
A. D.
Baczewski
,
T. D.
Beaudet
,
A.
Benali
,
M. C.
Bennett
,
M. A.
Berrill
,
N. S.
Blunt
,
E. J. L.
Borda
,
M.
Casula
,
D. M.
Ceperley
,
S.
Chiesa
,
B. K.
Clark
,
R. C.
Clay
,
K. T.
Delaney
,
M.
Dewing
,
K. P.
Esler
,
H.
Hao
,
O.
Heinonen
,
P. R. C.
Kent
,
J. T.
Krogel
,
I.
Kylänpää
,
Y. W.
Li
,
M. G.
Lopez
,
Y.
Luo
,
F. D.
Malone
,
R. M.
Martin
,
A.
Mathuriya
,
J.
McMinis
,
C. A.
Melton
,
L.
Mitas
,
M. A.
Morales
,
E.
Neuscamman
,
W. D.
Parker
,
S. D.
Pineda Flores
,
N. A.
Romero
,
B. M.
Rubenstein
,
J. A. R.
Shea
,
H.
Shin
,
L.
Shulenburger
,
A. F.
Tillack
,
J. P.
Townsend
,
N. M.
Tubman
,
B.
Van Der Goetz
,
J. E.
Vincent
,
D. C.
Yang
,
Y.
Yang
,
S.
Zhang
, and
L.
Zhao
, “
QMCPACK: An open source ab initio quantum Monte Carlo package for the electronic structure of atoms, molecules and solids
,”
J. Phys.: Condens. Matter
30
,
195901
(
2018
).
45.
M. A.
McGuire
,
H.
Dixit
,
V. R.
Cooper
, and
B. C.
Sales
, “
Coupling of crystal structure and magnetism in the layered, ferromagnetic insulator CrI3
,”
Chem. Mater.
27
,
612
(
2014
).
46.
J. L.
Lado
and
J.
Fernández-Rossier
, “
On the origin of anisotropy in two dimensional CrI3
,”
2D Mater.
4
,
035002
(
2017
).
47.
C.
Xu
,
J.
Fenj
,
H.
Xiang
, and
L.
Bellaiche
, “
Interplay between Kitaev interaction and single ion anisotropy in ferromagnetic CrI3 and CrGeTe3 monolayers
,”
npj Comput. Mater.
4
,
57
(
2018
).
48.
O.
Besbes
,
S.
Nikolaev
,
N.
Meskini
, and
I.
Solovyev
, “
Microscopic origin of ferromagnetism in the trihalides CrCl3 and CrI3
,”
Phys. Rev. B
99
,
104432
(
2019
).
49.
Y.
Zhang
,
X.
Wu
,
B.
Lyu
,
M.
Wu
,
S.
Zhao
,
J.
Chen
,
M.
Jia
,
C.
Zhang
,
L.
Wang
,
X.
Wang
,
Y.
Chen
,
J.
Mei
,
T.
Taniguchi
,
K.
Watanabe
,
H.
Yan
,
Q.
Liu
,
L.
Huang
,
Y.
Zhao
, and
M.
Huang
, “
Magnetic order-induced polarization anomaly of Raman scattering in 2D magnet CrI3
,”
Nano Lett.
20
,
729
(
2020
).
50.
L.
Webster
and
J. A.
Yan
, “
Strain-tunable magnetic anisotropy in monolayer CrCl3, CrBr3, and CrI3
,”
Phys. Rev. B
98
,
144411
(
2018
).
51.
W. B.
Zhang
,
Q.
Qu
,
P.
Zhu
, and
C. H.
Lam
, “
Robust intrinsic ferromagnetism and half semiconductivity in stable two-dimensional single-layer chromium trihalides
,”
J. Mater. Chem. C
3
,
12457
(
2015
).
52.
L.
Webster
,
L.
Liang
, and
J. A.
Yan
, “
Distinct spin-lattice and spin-phonon interactions in monolayer magnetic CrI3
,”
Phys. Rev. B
20
,
23546
(
2018
).
53.
M.
Cococcioni
and
S.
De Gironcoli
, “
Linear response approach to the calculation of the effective interaction parameters in the LDA+U method
,”
Phys. Rev. B
71
,
035105
(
2005
).
54.
G.
Kresse
and
J.
Hafner
, “
Ab initio molecular dynamics for liquid metals
,”
Phys. Rev. B
47
,
558
(
1993
).
55.
G.
Kresse
and
J.
Hafner
, “
Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium
,”
Phys. Rev. B
49
,
14251
(
1994
).
56.
G.
Kresse
and
J.
Furthmüller
, “
Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set
,”
Comput. Mater. Sci.
6
,
15
(
1996
).
57.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
,
11169
(
1996
).
58.
P. E.
Blöchl
and
J.
Furthmüller
, “
Projector augmented-wave method
,”
Phys. Rev. B
50
,
17953
(
1994
).
59.
G.
Kresse
and
D.
Joubert
, “
From ultrasoft pseudopotentials to the projector augmented-wave method
,”
Phys. Rev. B
59
,
1758
(
1999
).
60.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
, “
Restoring the density-gradient expansion for exchange in solids and surface
,”
Phys. Rev. Lett.
100
,
136406
(
2008
).
61.
J.
Krogel
, “
Nexus: A modular workflow management system for quantum simulation codes
,”
Comput. Phys. Commun.
198
,
154
(
2016
).
62.
M.
Burkatzki
,
C.
Filippi
, and
M.
Dolg
, “
Energy-consistent pseudopotentials for quantum Monte Carlo calculations
,”
J. Chem. Phys.
126
,
234105
(
2007
).
63.
T.
Ichibha
,
A. L.
Dzubak
,
J. T.
Krogel
,
V. R.
Cooper
, and
F. A.
Reboredo
, “
CrI3 revisited with a many-body ab initio theoretical approach
,”
Phys. Rev. Mater.
5
,
064006
(
2021
).
64.
C.
Lin
,
F.-H.
Zong
, and
D. M.
Ceperley
, “
Twist-averaged boundary conditions in continuum quantum Monte Carlo
,”
Phys. Rev. E
64
,
016702
(
2001
).
65.
M.
Holzmann
,
R. C.
Clay
,
M. A.
Morales
,
N. M.
Tubman
,
D. M.
Ceperley
, and
C.
Pierleoni
, “
Theory of finite size effects for electronic quantum Monte Carlo calculations of liquids and solids
,”
Phys. Rev. B
94
,
035126
(
2016
).
66.
M.
Casula
,
S.
Moroni
,
S.
Sorella
, and
C.
Filippi
, “
Size-consistent variational approaches to nonlocal pseudopotentials: Standard and lattice regularized diffusion Monte Carlo methods revisited
,”
J. Chem. Phys.
132
,
154113
(
2010
).
67.
A. L.
Dzubak
,
J. T.
Krogel
, and
F. A.
Reboredo
, “
Quantitative estimation of 3d transition metal pseudopotentials in diffusion Monte Carlo
,”
J. Chem. Phys.
147
,
024102
(
2017
).
68.
J.
Tiihonen
,
P. R. C.
Kent
, and
J. T.
Krogel
, “
Surrogate hessian accelerated structural optimization for stochastic electronic structure theories
” (unpublished) (
2021
).
69.
A.
Togo
and
I.
Tanaka
, “
First principles phonon calculations in materials science
,”
Scr. Mater.
108
,
1
5
(
2015
).
70.
B.
Yang
,
X.
Zhang
,
H.
Yang
,
X.
Han
, and
Y.
Yan
, “
Nonmetallic atoms induced magnetic anisotropy in monolayer chromium trihalides
,”
J. Phys. Chem. C
123
,
691
(
2019
).
71.
Z.
Wu
,
J.
Yu
, and
S.
Yuan
, “
Strain-tunable magnetic and electronic properties of monolayer CrI3
,”
Phys. Chem. Chem. Phys.
21
,
7750
(
2019
).
72.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
(
1996
).
73.
J. P.
Perdew
,
A.
Ruzsinszky
,
G. I.
Csonka
,
O. A.
Vydrov
,
G. E.
Scuseria
,
L. A.
Constantin
,
X.
Zhou
, and
K.
Burke
, “
Restoring the density-gradient expansion for exchange in solids and surfaces
,”
Phys. Rev. Lett.
100
,
136406
(
2008
).
74.
J.
Son
,
B. C.
Park
,
C. H.
Kim
,
H.
Cho
,
S. Y.
Kim
,
L. J.
Sandilands
,
C.
Sohn
,
J.-G.
Park
,
S. J.
Moon
, and
T. W.
Noh
, “
Unconventional spin-phonon coupling via the Dzyaloshinskii–Moriya interaction
,”
npj Quantum Mater.
4
,
17
(
2019
).
75.
D.
Kozlenko
,
O.
Lis
,
S.
Kichanov
,
E.
Lukin
,
N.
Belozerova
, and
B.
Savenko
, “
Spin-induced negative thermal expansion and spin–phonon coupling in van der Waals material CrBr3
,”
npj Quantum Mater.
6
,
19
(
2021
).
76.
A.
Zayak
,
P.
Entel
,
K.
Rabe
,
W.
Adeagbo
, and
M.
Acet
, “
Anomalous vibrational effects in nonmagnetic and magnetic Heusler alloys
,”
Phys. Rev. B
72
,
054113
(
2005
).
77.
B. H.
Zhang
,
Y. S.
Hou
,
Z.
Wang
, and
R. Q.
Wu
, “
First-principles studies of spin-phonon coupling in monolayer Cr2Ge2Te6
,”
Phys. Rev. B
100
,
224427
(
2019
).
78.
Z.
Liu
,
K.
Guo
,
G.
Hu
,
Z.
Shi
,
Y.
Li
,
L.
Zhang
,
H.
Chen
,
L.
Zhang
,
P.
Zhou
,
H.
Lu
 et al, “
Observation of nonreciprocal magnetophonon effect in nonencapsulated few-layered CrI3
,”
Sci. Adv.
6
,
eabc7628
(
2020
).
79.
D.-H.
Kim
,
K.
Kim
,
K.-T.
Ko
,
J.
Seo
,
J. S.
Kim
,
T.-H.
Jang
,
Y.
Kim
,
J.-Y.
Kim
,
S.-W.
Cheong
, and
J.-H.
Park
, “
Giant magnetic anisotropy induced by ligand LS coupling in layered Cr compounds
,”
Phys. Rev. Lett.
122
,
207201
(
2019
).
80.
I.
Lee
,
F. G.
Utermohlen
,
D.
Weber
,
K.
Hwang
,
C.
Zhang
,
J.
van Tol
,
J. E.
Goldberger
,
N.
Trivedi
, and
P. C.
Hammel
, “
Fundamental spin interactions underlying the magnetic anisotropy in the Kitaev ferromagnet CrI3
,”
Phys. Rev. Lett.
124
,
017201
(
2020
).
81.
J.
Cenker
,
B.
Huang
,
N.
Suri
,
P.
Thijssen
,
A.
Miller
,
T.
Song
,
T.
Taniguchi
,
K.
Watanabe
,
M. A.
McGuire
,
D.
Xiao
, and
X.
Xu
, “
Direct observation of two-dimensional magnons in atomically thin CrI3
,”
Nat. Phys.
17
,
20
25
(
2021
).
82.
A. V.
Chumak
,
V. I.
Vasyuchka
,
A. A.
Serga
, and
B.
Hillebrands
, “
Magnon spintronics
,”
Nat. Phys.
11
,
453
461
(
2015
).
83.
B.
Blaiszik
,
K.
Chard
,
J.
Pruyne
,
R.
Ananthakrishnan
,
S.
Tuecke
, and
I.
Foster
, “
The materials data facility: Data services to advance materials science research
,”
JOM
68
(
8
),
2045
2052
(
2016
).
84.
B.
Blaiszik
,
L.
Ward
,
M.
Schwarting
,
J.
Gaff
,
R.
Chard
,
D.
Pike
,
K.
Chard
, and
I.
Foster
, “
A data ecosystem to support machine learning in materials science
,”
MRS Communications
9
,
1125
1133
(
2019
).

Supplementary Material