Analog quantum chemistry simulation

1, 2 1, 3, 1, 4 5, 1 1, 6, Javier Arg¨uello-Luengo, Alejandro Gonz´alez-Tudela, ∗ Tao Shi, Peter Zoller, and J. Ignacio Cirac † 1Max-Planck-Institut f¨urQuantenoptik, Hans-Kopfermann-Straße 1, D-85748 Garching, Germany 2Institut de Ci`enciesFot`oniques(ICFO), The Barcelona Institute of Science and Technology, 08860 Castelldefels, Spain 3Instituto de F´ısica Fundamental IFF-CSIC, Calle Serrano 113b, Madrid 28006, Spain 4CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, P.O. Box 2735, 100190, China 5Center for Quantum Physics, University of Innsbruck, A-6020 Innsbruck, 6Munich Center for Quantum Science and Technology (MCQST), Munich, Germany (Dated: October 20, 2020) Computing the electronic structure of molecules with high precision is a central challenge in the field of quantum chemistry. Despite the enormous success of approximate methods, tackling this problem exactly with conventional computers is still a formidable task. This has triggered several theoretical [1,2] and experimental [3–5] efforts to use quantum computers to solve chemistry problems, with first proof-of-principle realizations done in a digital manner. An appealing alternative to the digital approach is analog quantum simulation, which does not require a scalable quantum computer, and has already been successfully applied in condensed matter physics problems [6–8]. However, all available or planned setups cannot be used in quantum chemistry problems, since it is not known how to engineer the required Coulomb interactions with them. Here, we present a new approach to the simulation of quantum chemistry problems in an analog way. Our method relies on the careful combination of two technologies: ultra-cold atoms in optical lattices and cavity QED. In the proposed simulator, fermionic atoms hopping in an optical potential play the role of electrons, additional optical potentials provide the nuclear attraction, and a single spin excitation over a Mott insulator mediates the electronic Coulomb repulsion with the help of a cavity mode. We also provide the operational conditions of the simulator and benchmark it with a simple molecule. Our work opens up the possibility of efficiently computing electronic structures of molecules with analog quantum simulation.

Quantum computers are expected to have a signifi- engineering long-range (Coulomb) interactions between cant impact in several areas of science, as they will be fermionic particles, and no system has been identified so able to tackle problems which are intractable with clas- far fulfilling such requirement. This is why current efforts sical devices. Particularly relevant are quantum many- concentrate in digital simulation. body problems involving several systems interacting with In this work we show how to build an analog simu- each other according to the rules of Quantum Physics [9]. lator for quantum chemistry problems by bridging two Among the most timely and significant ones are quantum paradigmatic systems, namely, ultra-cold atoms in opti- chemistry problems, which generally imply obtaining the cal lattices [15–17] and cavity QED [18–22]. Fermionic ground state energy of many electrons interacting with atoms trapped in a periodic 3D optical potential play both the nuclei and among themselves through Coulomb the role of electrons and are subject to additional op- interactions. Current approaches to the simulation of tical potentials emulating their interaction with the nu- chemistry problems with quantum computers follow the clei. The key ingredient of the scheme is to trap another digital approach [10–13], in which one breaks the com- atomic species in the Mott insulator regime, with sev- plete Hamiltonian in gates that are applied in a time eral internal states such that its spin excitations medi- dependent manner. ate effective forces between the simulated electrons. We An alternative way to address quantum many-body show that even though the interaction is local, one can problems is analog quantum simulation [14]. The idea is arXiv:1807.09228v2 [quant-ph] 17 Oct 2020 induce Coulomb-like forces among the fermionic atoms to use a well-controlled quantum system (the simulator) in a scalable manner. While the setup is discrete and and engineer its interactions according to the Hamilto- finite, we show that precise results can be obtained for nian under investigation. This approach has already ad- simple molecules with moderate lattice sizes. Apart from dressed questions that the most advanced classical com- the standard advantages of analog simulation over quan- puters cannot compute [6–8]. The key feature is that tum computing regarding the required control [14], the their interactions are either local or short-range, being present scheme does not rely on a judicious choice of ideally suited for the existing simulators. On the con- molecular orbitals [23], but directly operates in real space trary, analog simulation of quantum chemistry requires improving convergence to the exact result as the system size increases. One of the main goals of Quantum Chemistry is to ob- tain the low energy behavior of Ne electrons and several ∗ [email protected] nuclei when the positions, rn, of the nuclei are fixed. Us- † [email protected] ing a cubic discretization in real space of N N N sites, × × 2

a Electronic molecular Hamiltonian b

Kinetic Nuclear Electronic + + term attraction repulsion b | i

tF a Jc | i U V0 a

Zn V0 tF fi†fj V0 fj†fj + fi†fifj†fj j rn i j i,j j,n i ,j k k hXi X k k X

c

J

g

trapped 0 Mottspin fermion | i excitationexcitation Fermionic atom Mediating atom

Fig. 1 | Schematic representation of the analog simulator. a, Fermionic atoms, playing the role of electrons, are trapped in a periodic 3D cubic potential. Their hopping simulates the kinetic energy term of the electrons, and they are subject to additional optical potentials which emulate the nuclear interaction. b, Coulomb repulsion among the fermions is mediated by a spin excitation of a Mott insulator with three internal levels. Excitations in level |bi are allowed to propagate through spin exchange interactions with rate J (upper in the box). Excitations in level |ai experience a strong repulsive interaction with the fermions and interacts with a cavity mode. These two levels are coupled either through a microwave or a two-photon Raman transition (lower in the box). c, Illustration of the complete simulator for the H2 molecule. While a bidimensional lattice is pictured, the experimental proposal presented here refers to a three-dimensional optical lattice.

the electronic Hamiltonian to solve contains three terms, where a0, a and Ry are the Bohr radius, lattice spacing, Hqc = Hkin + Hnuc + He e, (using ~ = 1, and dropping and Rydberg energy, respectively. Thus, we work in a the spin index) − regime,

1/3 (a) 1 2tF /V0 N/Ne , Hkin = tF f †fj , (1)   − i i, j hXi such that the first inequality prevents discretization ef- fects, and the second guarantees the molecule fits in the H = Z V ( j r ) f †f (2) nuc − n k − nk j j volume of the simulator. n, j X Our simulator then requires three ingredients (see He e = V ( i j ) f †fif †fj , (3) Fig. 1a): i) cold spin-polarized fermionic atoms hopping − k − k i j Xi, j in a 3D optical potential with a tunable tunneling rate, JF , to play the role of electrons [15]. We consider spin- where fi are annihilation operators of electrons at site i less fermions; the spin degree of freedom can be included fulfilling fi, f † = δi,j, and i, j denote nearest neigh- using an extra internal level [24] (see Methods). ii) Addi- { j } h i bour sites. Hkin describes the electrons hopping at a rate tional potentials to emulate the nuclei attraction. Since tF , Hnuc represents the nuclear attraction when the nu- this is a single-particle Hamiltonian, it can be created clei are at positions rn, while He e accounts for electron- through optical Stark-Shifts with an adequate spatial − electron repulsion. In both cases, the attractive/repulsive modulation. For example, one can use holographic tech- potential has the standard Coulomb form, V (r) = V0/r. niques [25, 26] with judiciously optimized phase masks to The connection of the length/energy scales of the discrete engineer a Coulomb-like spatial potential at the fermionic Hamiltonian Hqc and the continuum one is given by: positions (see Methods). iii) The most difficult part is to simulate He e, since it involves repulsive interactions 2 − a0/a = 2tF /V0 and Ry = V0 /(4tF ) , (4) between the fermions with a 1/r dependence. Inspired 3

by how virtual photons mediate Coulomb interactions in 1.0 0 10 QED, we use a spin excitation of another atomic species a b |/Ry 0 /Ry -2 forming a Mott insulator to mediate the Coulomb forces 0 10 E between fermions (see Fig. 1b). It is composed of N 3 |ΔE M 0 0 atoms trapped in an optical potential with the same spac- 0.0 10 4·10 ing as the fermions, and with two additional internal atomic states, a and b , which describe spin excita- tions. Spin excitations| i in| ai state interact repulsively and locally with the fermionic atoms, with strength U, and -1.0 a0/a<1 propagate through the long-range couplings induced by 100 101 a cavity mode, with rate Jc [18–22]. The b internal tF/V0 c state is subject to a different optical potential,| i such that its itinerant spin excitation propagates through standard nearest-neighbor exchange [27], at rate J. Furthermore, an external field (Raman or a RF field) drives the tF/V0 = 0.3 tF/V0 = 1 tF/V0 = 6 a b transition with coupling strength g, and detuning |∆.i−| Thei complete simulator Hamiltonian after the elimi- Fig. 2 | Atomic Hydrogen spectrum dependence on the effective Bohr radius. a, Lower part of the spectrum nation of the cavity mode reads Hsim = Hkin+Hnuc+HM , with of the atomic Hamiltonian Hqc for a cubic lattice of N = 100. Dashed lines indicate the position of the 3 first atomic levels 3 associated to the continuum Hamiltonian. In the colored re- HM = ∆ b†bj + J b†bj + Jc/N a†aj j i M i gion, tF /V0 < 0.5, the Bohr radius is smaller than the lattice j i,j i,j X Xh i X (5) spacing and energies are highly affected by the cut-off of the nuclear potential. As the hopping parameter tF /V0 increases, + U a†a f †f + g a†b + b†a , j j j j j j j j the system is effectively zoomed-in, as we show in the panels Xj Xj   for the electron density of the second-lowest energy orbital. This includes more lattice sites in the simulation, reducing being aj/bj annihilation operators for a a / b -spin exci- −2 | i | i systematic deviations as (tF /V0) (red dashed line), as we tation in site j. Intuitively, the on-site interaction U lo- show in the inset for the lowest energy state (see Supple- calizes the a / b excitations around the fermions, renor- mentary Information section 1). At higher values of t /V , | i | i F 0 malizing their tunneling rates and creating an effective solutions suffer from finite-size effects. b, Axial cut in the interaction. Mathematically, one can adiabatically elim- central positions of the lattice is represented for the first 9 inate the Mott insulator excitations and derive the effec- eigenstates of Hqc for tF /V0 = 2, N = 150. c, Appropriately tive dynamics for the fermions (see Supplementary Infor- choosing the Bohr radius, we show how the same orbital can mation section 2). The fermionic part of the simulator be obtained with N = 1000 (up, tF /V0 = 150) or N = 20 Hamiltonian H H , with t = J (N 1)/N , (down, tF /V0 = 3), where the discretization of the system is sym qc F F e e more noticeable. where the electron-electron≈ potential follows a− Yukawa form [28] with a constant energy shift C: must be larger than the fermionic lattice N, but smaller V0 r/L V (r) C + e− , (6) than the Mott insulator size, that is: ≈ r/a

(e) N L/a < NM . where L/a = J/(U + ∆ + ρM Jc 6J) is the localiza-  − g2 tion length, which can be tuned with ∆, and V0 = When all (a-e) inequalities are satisfied, the exact so- p 2πJNe the strength of the potential repulsion. Here ρM = lution in the continuum limit is recovered in the limit 3 NM > N . Thus, the finite size of the simulator is Ne/NM . This mapping between Hsym and Hqc holds → ∞ as long as, what ultimately limits the precision of the simulation. We now benchmark our simulator for moderate system (b) Jc U, sizes using numerical simulations. In Fig. 2 we solve the  Hydrogen problem in a lattice to explore discretization (c) J J ρ N , and V N J ρ , F  c M e 0 e  c M and finite size effects by comparing the energies of the low 2 2 (d) V 0pNe JN (a/Lp ) . excited states with that of the continuum. We show that  an error of 0.3% with respect to the exact energy can be Condition (b) enforces that the a excitation localizes obtained for systems of N = 100. In Fig. 3 we analyze symmetrically only around the position of the fermions; the accuracy for the simplest molecule, H2. First, we (c) guarantees that neither the tunneling of the fermions compute exactly the energy of the spin excitation that nor the interaction with the b-excitations spoils the effec- mediates the fermionic repulsion, as a function of the tive interaction; (d) ensures that the Yukawa potential interfermionic separation (Fig. 3a). We show that it re- does not depend on the fermionic positions. Further- produces the 1/r behavior over a wide range of values of more, to obtain a truly Coulomb repulsion, the length L g/J and L. In Fig. 3b we compute the molecular poten- 4

10-7 -1.6 a A -2.3 b B -C) /J

E / Ry -2.4 ( V -9 10 1 2 -2.0

10-11 g=10-3J, L=100a L= ∞ L=1000a g=10-4J, L=100a L=350a g=10-4J, L=35a -2.4 100 101 102 0 2 4 6 r/a d/a0 10-1 c d L= ∞ 0 L=1000a -2.0 L=350a -C) /V ( V 100 E / Ry L=35a -2.5 -2 10 F=1 F=1 10-2 F=1.5 F=1.5 1 2 L=35a 10 10 -3.0 101 102 0 2 4 6 r/a d/a0

Fig. 3 | Molecular potential and effective interaction mediated by the Mott insulator. a, Energy of the single- excitation bound state of Hamiltonian (5) for two fixed fermions as a function of their separation, r. We choose ∆ = 2J, NM = 200,Jc = J, such that (a-e) inequalities are satisfied. The Yukawa potential of Eq. (6) corresponding to each configuration of parameters is plotted with dashed lines. b, We use this effective interaction to calculate the molecular potential associated to an analog simulator of H2 of size N = 75. For each internuclear separation, we choose tF /V0 giving optimal accuracy (see Supplementary Information section 3), ranging from tF /V0 = 4.2 to tF /V0 = 2.3 in the dissociation limit (dotted line). Molecular orbitals are included in the projective basis until convergence is observed. For a Coulomb potential (blue dots), the result agrees with an accurate solution in the nonrelativistic regime [29, 30] (dashed line). As L decreases, the exponential decay in the Yukawa potential prevails, underestimating Coulomb repulsion and lowering the molecular potential. c, This underestimation of the repulsive potential is stressed when violating N  L/a (see inset). d, Changing the ratio, F between the electronic and nuclear potential, one can explore artificial repulsive interactions that form pseudomolecules in more relaxed experimental conditions. The dotted line here represents the limit of zero-repulsion in the absence of a mediating excitation. tial with N = 75 by using a Yukawa electronic potential peating the measurements. In the Methods section we with different lengths L. We observe excellent qualita- also analyze some sources of errors, and ways to circum- tive agreement for all L’s considered in the figure, and a vent them. We emphasize that some of the elements quantitative matching when L aN. Remarkably, even and conditions required in this proposal are beyond the if L < aN, valuable information can still be extracted state of the art. However, the rapid progress of analog by adjusting∼ other experimental parameters. In Fig. 3d, quantum simulation may well lead to the realization of we illustrate how one can increase the V0 of the electron the present ideas in the near future, motivated by its repulsion to compensate the underestimation of the po- potential impact in the determination of chemical struc- tential at long distances, and obtain a pseudomolecular tures, the understanding of reaction mechanisms, or the potential that is qualitatively similar to the one expected development of molecular electronics. Furthermore, with with Coulomb interactions. judicious changes in the implementation, e.g., different We consider some practical issues for the implemen- optical potential geometries, conditions (b-d) may be re- tation of these ideas (see Methods). As atomic species laxed. Thus, we believe our proposal is a promising com- fulfilling the requirements we propose to use two iso- plementary approach to the fault-tolerant quantum com- topes of alkali-earth atoms, namely 87Sr and 88Sr for the puter required to solve the same problems. fermions and bosons, respectively. The quantum simu- In summary, we have shown how to simulate quantum lator can be initialized by using adiabatic preparation, chemistry problems using cold atoms in optical lattices where the hopping, nuclear attraction, and interactions embedded in a cavity. We expect the present proposal are sequentially turned on. We also propose to read to stimulate both theoretical and experimental research, out the physical properties by measuring the total en- even before the realization of a fully-fledged analog simu- ergy of the system, by detecting the kinetic energy of lator for quantum chemistry. Fig. 4 provides a roadmap the fermionic atoms under different conditions and re- of experiments with increasing complexity (as moving to 5

Nuclear Number of Electronic Dimension potential electrons repulsion the bottom right of the table) towards complete ana- log chemistry simulation. For instance, first experiments a 1D d Focalized f 1 atom i Single boson could be performed with spin-less fermions in one and two laser trapped fermion spatial dimensions. Another simplification might come from non-Coulomb nuclear potentials, e.g., in form of a b 2D g 2 atoms j Mott insulator trapped gaussian, which does not require holographic techniques; fermiontrapped fermion or by using simpler schemes to obtain the fermionic po- Experimentalpathway tential, e.g., using a single boson instead of a Mott in- Holographic c 3D e h +2 atoms k Mott + cavity potential sulator, and without a cavity. The latter still provides

trapped trapped an effective repulsion between two fermions. With these fermiontrapped fermion trapped fermion simplifications, there is a clear pathway from state-of-the- Chemistry fermion… art setups towards more challenging experimental setups based on progress in technology. Most important, in all these intermediate proof-of-principle setups one could al- Fig. 4 | Experimental pathway. Schematic representation ready observe molecular-like potentials, dissociation, and of the simplifications that can be considered for the different other basic phenomena in chemistry. Besides, such ex- interactions of the system. The lowest line c+e+h+k corre- periments can also prove valuable to benchmark various sponds to the full proposal in this work, leading to the chem- ical interactions we observe in nature. Chemistry in lower numerical techniques, and trigger the development of new dimensions could be considered by restricting the optical lat- theoretical methods, and thus reaching a deeper under- tice in such directions (see a+b). The holographically created standing of the problems that appear in chemistry and Coulomb potential could be replaced by, e.g., the Gaussian which are challenging to test with classical computers. profile of a focalized laser (see d), giving a different scaling for electron-nuclei attraction. First implementations with sin- gle atoms (see f) would allow to observe simple electronic or- bitals (such as the energy levels of Hydrogen) with no need to mediate repulsion. For only two atoms (see g) one doesn’t need the symmetrizing effect of the cavity when mediating the Coulomb repulsion with the Mott insulator (see j). Different scalings for repulsion can also be explored in more simplified setups, such as using a single boson that hops on the lattice and has on-site interaction with the fermionic atoms (see i).

[1] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, M. Head- [10] Y. Cao, et al., Chem. Rev. 119, 10856-10915 (2019). Gordon, Simulated Quantum Computation of Molecular [11] S. McArdle, S. Endo, A. Aspuru-Guzik, S. Benjamin, Energies, Science 309, 1704-1707 (2005). X. Yuan, Quantum computational chemistry, Rev. Mod. [2] D. Wecker, B. Bauer, B. K. Clark, M. B. Hastings, Phys. 92, 015003 (2020). M. Troyer, Gate-count estimates for performing quan- [12] I. Kassal, J. D. Whitfield, A. Perdomo-Ortiz, M.-H. tum chemistry on small quantum computers, Phys. Rev. Yung, A. Aspuru-Guzik, Simulating chemistry using A 90, 022305 (2014). quantum computers, Annu. Rev. Phys. Chem 62, 185- [3] P. O’Malley, et al., Scalable Quantum Simulation of 207 (2011). Molecular Energies, Phys. Rev. X 6, 031007 (2016). [13] M. Reiher, N. Wiebe, K. M. Svore, D. Wecker, M. Troyer, [4] A. Kandala, et al., Hardware-efficient variational quan- Elucidating reaction mechanisms on quantum computers, tum eigensolver for small molecules and quantum mag- Proc. Natl. Acad. Sci. U. S. A. 114, 7555-7560 (2017). nets, Nature 549, 242-246 (2017). [14] J. I. Cirac, P. Zoller, Goals and opportunities in quantum [5] B. P. Lanyon, et al., Towards quantum chemistry on a simulation, Nat. Phys. 8, 264-266 (2012). quantum computer, Nat. Chem. 2, 106-111 (2010). [15] I. Bloch, J. Dalibard, W. Zwerger, Many-body physics [6] J.-y. Choi, et al., Exploring the many-body localization with ultracold gases, Rev. Mod. Phys. 80, 885-964 (2008). transition in two dimensions., Science 352, 1547-1552 [16] K. W. Murch, K. L. Moore, S. Gupta, D. M. Stamper- (2016). Kurn, Observation of quantum-measurement backaction [7] S. Trotzky, et al., Probing the relaxation towards equilib- with an ultracold atomic gas, Nat. Phys. 4, 561-564 rium in an isolated strongly correlated one-dimensional (2008). Bose gas, Nat. Phys. 8, 325-330 (2012). [17] L. Krinner, M. Stewart, A. Pazmi˜no, J. Kwon, [8] M. Greiner, O. Mandel, T. Esslinger, T. W. H¨ansch, D. Schneble, Spontaneous emission of matter waves from I. Bloch, Quantum phase transition from a superfluid to a tunable open quantum system, Nature 559, 589-592 a Mott insulator in a gas of ultracold atoms, Nature 415, (2018). 39-44 (2002). [18] H. Ritsch, P. Domokos, F. Brennecke, T. Esslinger, Cold [9] R. P. Feynman, Simulating Physics with Computers, Int. atoms in cavity-generated dynamical optical potentials, J. Theor. Phys. 217 (1982). Rev. Mod. Phys. 85, 553-601 (2013). 6

[19] S. Schreppler, et al., Quantum metrology. Optically mea- AUTHOR CONTRIBUTIONS suring force near the standard quantum limit, Science 344, 1486-1489 (2014). J.A.-L., A.G.-T. and T.S. designed the model, devel- [20] F. Brennecke, S. Ritter, T. Donner, T. Esslinger, Cavity oped the methods, and analyzed the data. J.A.-L. and optomechanics with a Bose-Einstein condensate, Science A.G.-T. performed the calculations. J.A.-L., A.G.-T. and 322, 235-238 (2008). J.I.C. wrote the manuscript with input from all authors. [21] F. Brennecke, et al., Cavity QED with a Bose–Einstein J.I.C. and P.Z. conceived the study and were in charge condensate, Nature 450, 268-271 (2007). of overall direction and planning. [22] P. Domokos, H. Ritsch, Collective Cooling and Self- Organization of Atoms in a Cavity, Phys. Rev. Lett. 89, 253003 (2002). [23] A. Szabo, N. S. Ostlund, Modern quantum chem- METHODS istry: introduction to advanced electronic structure theory (Courier Corporation, 2012). Here we provide further details and candidates for the [24] A. J. Daley, M. M. Boyd, J. Ye, P. Zoller, Quantum com- experimental implementation of some of the key ingredi- puting with alkaline-earth-metal atoms, Phys. Rev. Lett. 101, 170504 (2008). ents of this proposal. [25] M. R. Sturm, M. Schlosser, R. Walser, G. Birkl, Quan- tum simulators by design: Many-body physics in recon- figurable arrays of tunnel-coupled traps, Phys. Rev. A Holographic engineering of the nuclear potential 95, 063625 (2017). [26] D. Barredo, V. Lienhard, S. de L´es´eleuc, T. Lahaye, A. Browaeys, Synthetic three-dimensional atomic struc- Holographic techniques are a candidate approach to tures assembled atom by atom, Nature 561, 79-82 (2018). engineer the three-dimensional Coulomb potential seen [27] L. Riegger, et al., Localized Magnetic Moments with by the fermionic atoms, Hnuc. Following this method, Tunable Spin Exchange in a Gas of Ultracold Fermions, the potential is created by imprinting a phase pattern on Phys. Rev. Lett. 120, 143601 (2018). an incoming laser, as used in Ref. [26]. There, it allows [28] I. de Vega, D. Porras, J. I. Cirac, Matter-Wave Emis- to experimentally induce microtrap potentials for single sion in Optical Lattices: Single Particle and Collective atoms, and the same technique can engineer more general Effects, Phys. Rev. Lett. 101, 260404 (2008). intensity patterns at the lattice sites where the fermions [29] W. Kolost, L. Wolniewlczt, Accurate Adiabatic Treat- are trapped by choosing an appropriate phase mask. For ment of the Ground State of the Hydrogen Molecule, J. this, we have particularized the algorithm used in that Chem. Phys. 41, 3663-3673 (1964). [30] J. S. Sims, S. A. Hagstrom, High precision variational implementation (Gerchberg-Saxton algorithm [31], G-S) calculations for the Born-Oppenheimer energies of the to identify the phase mask leading to the particular ge- ground state of the hydrogen molecule, J. Chem. Phys. ometry needed in our proposal (a 3D Coulomb potential, 124, 094101 (2006). Vj = V0/ j n ). k − k We then use the ping-pong strategy described by the discretized G-S algorithm to retrieve the phases of the ACKNOWLEDGMENTS holographic field u associated to our intensity distribu- 2 tion ( uj = Vj). This approach iterates over the fol- We thank S. Blatt and N. Meyer for useful discus- lowing| steps| [32]: (i) The field is transformed to recip- sions regarding the experimental implementation of this rocal space, obtainingu ˜ = FFT(u). (ii) We restrict to proposal. The authors acknowledge the ERC Advanced those k-vectors that satisfy the physical constraints, i.e. Grant QENOCOBA under the EU Horizon 2020 pro- they sit on the Ewald sphere of radius k0 = 2π/λ, be- gram (grant agreement 742102). J.A.-L. acknowledges ing λ the wavelength of the monochromatic input light. support from ”la Caixa” Foundation (ID 100010434) un- The rest of contributions are neglected, defining a con- der the fellowship LCF/BQ/ES18/11670016. Work at strained field in Fourier space,u ˜c. (iii) We transform ICFO is supported from the Spanish Ministry of Econ- back to real space, uc = IFFT(˜uc). At this point, uc omy and Competitiveness through the ”Severo Ochoa” satisfies diffraction laws, but may differ from the goal program, Fundaci´oPrivada Cellex, and from Generalitat potential. (iv) The objective potential is combined with de Catalunya through the CERCA program. A. G.-T. ac- the phases of the constrained solution, ϕ = arg uc , knowledges support from the Spanish project PGC2018- j j iϕj 094792-B-100 (MCIU/AEI/FEDER, EU) and from the giving uj = Vje . To improve the accuracy, we use a CSIC Research Platform on Quantum Technologies PTI- refined phase mask of size (ndiv N) (ndiv N), being p · × · 001. T.S. is grateful to the Thousand-Youth-Talent Pro- N the number of lattice sites per side, and ndiv a refin- gram of China. P.Z. was supported by the ERC Synergy ing factor (see insets in the Extended Data Fig. 1). As Grant UQUAM, and SFB FoQus by the Austrian Science the trapped atoms are only affected by the value of the Fundation. holographic potential on the lattice sites, these are the only coordinates where the field is updated in step (iv). In the Extended Data Fig. 1 we show the result of applying such optimization process to the Coulomb po- 7

1 tential. By quantifying a normalized relative error [33], S0 level off resonance can be used to induce the exter- nal potential (with the holographic techniques explained 1 2 above). Note that the isotope shift is sufficiently large  = V V / V , (7) N 3 j∗ − j | j| in those atoms so that this can be negligible even in the j 1 X q p  case that the level S0 is chosen as 0 . The cavity can 3 3 | i be tuned close to the P0 S1 resonance, without af- between the desired Coulomb potential (Vj) and the in- → 2 fecting the other states. Two-photon Raman transitions tensity pattern numerically obtained (V ∗ = uj ), we 3 3 j | | can then couple the levels P0 and P2 appropriately, or observe that the accuracy obtained for ndiv = 3 already 1 one can also use a four-photon transition using level S0. provides normalized relative errors smaller than 0.3% for Also, note that having nuclear spin 0 is not a problem N = 30. More precise results can be obtained by in- in this case, as the mediated electronic repulsion is spin- creasing the iterations of the algorithm or the refining independent. By choosing an isotope of Sr, an optical factor. lattice with the same spacing can be engineered at the 1 bosonic state S0, with similar choices to the fermionic atoms. Furthermore, the lattice depth and spacing in 3 Candidate atomic species P2 can be independently controlled by, e.g., using magic wavelengths [34], and its scattering length can be tuned This proposal is based on the interplay between two using, e.g., non-resonant light [35]. atomic species: (i) Fermionic atoms: they have two in- ternal levels, that play the role of the electronic spin. (ii) Mediating atoms: they must have three levels available: Adiabatic preparation 0 for the ground state in the Mott insulator; a with the| i spin excitation; b with the state that tunnels| i and A crucial step is the preparation of the appropriate | i induces the effective repulsion. states of Hsim, which needs to be carefully designed for All atoms need to see the same lattice, although with the molecular system one wants to explore. In general, tunable tunneling amplitudes. The fermions must ad- one is interested in the ground state properties of the ditionally see the external potential and have to interact electronic configuration for given positions of the nuclei. with the internal state of the mediating atoms. The scat- In this line, one option consists on sequentially adding tering lenghts for the interactions corresponding to levels the different interactions involved in the simulation. 0 , a and b do not need to be the same, and we re- As a particular example, let us illustrate this approach | i | i | i quire that the scattering length between the fermions and for the ground state of H2. It corresponds to a singlet, the mediating atom in state b is neglectful, so that the electrons then have opposite spin, and are therefore dis- scaling of the moving excitation| i is unchanged. Addition- tinguishable. We propose the following strategy. First, ally, the mediating atoms in a have to be exposed to we create a product state of tightly trapped fermions the cavity mode. Also, note that| i the atoms that mediate that respects the symmetries of the problem. Second, the fermionic interactions can correspond to both bosons we slowly allow fermions to hop. The attraction to the and fermions, as there is just a single excitation. nuclear potential then leads to single particle orbitals. Over the last years, many atomic species have been Finally, we adiabatically introduce the electronic repul- trapped, condensed, and used in experiments with op- sion, completing the entire Hamiltonian. tical lattices. For illustration, let us give a particular As an initial state, we begin with both fermions tightly example based on fermionic and bosonic alkaline-earth trapped in a lattice site j0 (tF (0) = 0). Using an exter- atoms. These atoms offer a rich internal structure, with nal laser we create a single Mott excitation in level a 3 3 at this position. The initial on-site interaction U = 0 long-lived excited metastable states P0 and P2. 6 As a particular choice, we can use 87Sr as a simula- with the fermionic atoms maintains this excitations at j . We can define this state as ψ = f † f † a† 0 , tor for electrons. Its nuclear spin is I = 9/2 and, sim- 0 0 ,j0 ,j0 j0 | i ↑ ↓ | i ilarly to [24], one can encode the spin of the simulated where f †/ ,n denotes the creations operator of one spin 1 ↑ ↓ electron into nuclear states, S0, mI = 9/2 and |↑i ≡ − up/down fermionic atom in position n, and an† the cre- 1S , m = 7/2 . This information is therefore |↓i ≡ 0 I − ation of a single excitation of the Mott insulator in that protected from the electronic transitions used in the rest site. One can now adiabatically evolve this initial state of the proposal. according to the following steps. One can now use one of its bosonic isotopes, 88Sr for 1 the mediating atoms. We assign the long-lived states S0, 1. Both the fermionic atoms and the Mott excitations 3P , and 3P as levels 0 , a , and b , respectively. Since are coherently transported into a symmetric super- 0 2 | i | i | i there exists a magic wavelength that makes the trapping position of two positions, n1/2, where the nuclear of the states S and one of the P equal, it is possible to potentials will be centered, separated by the desired choose nearby frequencies that provide basically the same number of sites d/a. This is, a new state ψI = 3/2 | i † † † † lattice period for all states with the appropriate condi- 2− (f ,n1 +f ,n2 )(f ,n1 +f ,n2 )(an† 1 +an† 2 ) 0 . One tions. Apart from that, an additional laser driving the can use↑ different↑ strategies↓ ↓ for doing this,| suchi as 8

using moving double well potential that adiabati- 3D spatial tomography of the Ne electron and reconstruct cally transports the wavepacket in opposite direc- the fermionic density. This is very complex in practice, tions [36], or using standing waves [37]. As the but with gas microscope techniques (see Ref. [39], for on-site interaction (U) is present, the Mott excita- instance) could be feasible. tion is also transported, mediated by the long-range An alternative would be measuring the energy of the cavity interaction JC . system. In addition to constructing molecular potentials, At this point, the holographic Coulomb potential scanning the energy at different nuclear configurations described in the previous sections is induced. Given can provide additional information, such as the value of that hopping processes are inhibited and fermionic molecular forces (Hellmann-Feynman theorem [23]). For atoms are now already centered in the nuclear po- this, three elements need to be simultaneously measured: the kinetic energy K , the nuclear attraction V , and sitions, no evolution will be observed. Also, the h i h ni coupling between the excited levels a/b is switched- the electronic repulsion Ve ; such that the total energy writes as E = K + V h+iV . Using sudden quenches off (g = 0). Thus, the repulsive interaction medi- h i h ni h ei ated by the Mott insulator is inactive. The on- of the Hamiltonian, such contributions can be indepen- site interaction U modifies the fermionic hopping dently converted into kinetic energy. One can then per- form a time of flight measurement of the fermionic atoms (JF tF ) as detailed in the main text. Given that → expelled from the lattice, using for example ionization or the optical lattice is infinitely deep (tF = 0) at this point, it translate into an effective null Bohr radius fluorescence techniques. As one can observe, the mea- sured quantities will not correspond to eigenstates of the (a0(0)/a = 0). The nuclear separation is then in- original Hamiltonian and this will give some variance in finite (d/a0 = ), and we have therefore prepared the dissociation∞ limit. the measurement proportional to the number of fermions. One could then repeat this procedure to gain statistically 2. The next step consists on increasing the orbital size. significance. Once the equilibrium point of the molecular For this, we adiabatically relax the optical lattice, potential is identified, the procedure can also be highly slowly increasing the value of tF . From the point of simplified as only measuring K is needed to read the view of the molecular potential, this corresponds to total energy at that point, basedh i on the virial theorem growing the Bohr radius, therefore decreasing the for molecules [40]. effective distance d/a0 (even though d/a remains fixed). As only the kinetic and attractive terms of the Hamiltonian are present, there is no interaction Experimental considerations between the two fermions, and the resulting eigen- state corresponds to two independent ground state A reliable simulation of the quantum chemistry Hamil- + electrons associated to H2 . tonian needs that our simulator, described by Eq. 5 in the main text, satisfies a set of inequalities (a-e). We are, 3. Once these single particle orbitals are prepared, however, aware that there will be other experimental im- one can adiabatically introduce the repulsive in- perfections that may impose extra conditions and that teractions. For this, we slowly couple Mott levels will have to be analyzed in detail to optimize the perfor- a and b using a Raman transition of intensity | i | i mance of the simulation. Among the more relevant ones 0 g. As we derived in the manuscript, this is the are: final→ ingredient required to induce an effective re- pulsion among the fermionic atoms, proportional to • Finite temperature leads to thermal fluctuations g2. One can then increase g until repulsion equals which may spoil the simulation. Thus, these fluctu- V0. ations will lead to defects in the Mott insulator (see below), an may also influence the internal states of In the Extended Data Table I we have summarized the atoms. The latter, however, can typically be the parameters that are modified at each stage. In the very well controlled in atomic systems as we just Extended Data Fig. 2 we numerically simulate this adi- need the atoms to be initially in a polarized state, abatic path in 2D (and 1/r-interactions) to prepare the which is reasonably easy to prepare. ground state for a nuclear separation d/a = 10, showing 0 Dephasing that such adiabatic preparation is indeed feasible within • can be originated by fluctuations in the reasonable parameters. transitions or due to magnetic fields (as internal levels are being used). This would spoil the poten- tial of the system as a . How- Measurement ever, the first effect is small in the case of microwave or Raman transitions, and the second can be con- trolled in the conditions already used for condensed From the chemistry perspective, all relevant quanti- matter simulations [6,7]. ties can be expressed in terms of the fermionic density. This is for example the approach used in DFT meth- • Inexact fermionic filling. Since fermions play the ods [43]. One possibility then consists on performing a role of electrons, an inexact number of fermionic 9

atoms hopping in the lattice translates into an er- The fluctuations of fi,j around 1 will induce an roneous effective charge in the simulated molecule. extra decoherence timescale, Γc,inh, that must be These errors can be possibly post-selected by mea- smaller than our simulator parameters as well. suring the number of fermionic atoms after the sim- In state-of-the-art experiments, optical cavities at ulation is performed. wavelengths of 780 nm and beam-waist of 60 µm are already available [41], which would roughly∼ al- • Defects in the Mott insulator. The absence of Mott low for 50 100 local minima of the standing wave particles in a given lattice site will locally mod- sitting in a− homogeneous region. ify the effective fermion potential. Fermions hop- ping to this site cannot mediate its repulsive in- • Cavity & atom losses. Even though the cavity- teraction through spin-excitations, perturbing the mediated interactions are mediated by a virtual simulated molecular orbital around this position. population of photons, the cavity decay introduces Importantly, the defects will not affect the poten- extra decoherence into the system due to the emis- tial far from the fermion such that the final perfor- sion of these virtual photons. Moreover, the atomic mance of the simulation will scale with the density excited states, also virtually populated, may as well of defects rather than their number. decay to free space introducing losses. Thus, the cooperativity of the cavity QED system must be • Spatial inhomogeneities of cavity coupling. In the large to avoid both type of losses. simulator Hamiltonian of Eq. (5) of the main text, we have assumed that the a -excitations couple ho- • Three-body losses. Since we have fermions and mogeneously to the cavity| mode.i In general, there there can be at most one atom per lattice site, these might be some inhomogeneities that translates in a type of losses should be small. Hamiltonian: From these qualitative arguments, we see that most of the possible errors of the simulation are either already under control in current experiments [6,7] or scale in an Jc f a†a . (8) intensive way. N 3 i,j i j M i,j X

[31] R. Gerchberg, O. Saxton, A practical algorithm for the Van Camp, V. Vuleti´c,Optomechanical Cavity Cooling determination of phase from image and diffraction plane of an Atomic Ensemble, Phys. Rev. Lett. 107, 143005 pictures, Optik 35, 237-246 (1971). (2011). [32] H. Chen, Y. Guo, Z. Chen, Experimental demonstration of holographic three-dimensional light shaping using a Gerchberg-Saxton algorithm, New J. Phys 7, 117 (2005). COMPETING INTERESTS [33] V. A. Soifer, V. Kotlar, L. Doskolovich, Iteractive Meth- ods For Diffractive Optical Elements Computation (CRC The authors declare no competing interests. Press, 2014). [34] J. Ye, H. Kimble, H. Katori, Quantum state engineer- ing and precision metrology using state-insensitive light CODE AND DATA AVAILABILITY traps, Science 320, 1734-1738 (2008). [35] A. Crubellier, R. Gonz´alez-F´erez,C. P. Koch, E. Luc- The computer code developed for this Letter is avail- Koenig, Controlling the s-wave scattering length with able upon reasonable request to the corresponding au- nonresonant light: Predictions of an asymptotic model, thors. All data supporting the findings of this study Phys. Rev. A 95, 023405 (2017). can be generated using the numerical methods described [36] O. Mandel, et al., Coherent transport of neutral atoms within the Methods and Supplementary Information. in spin-dependent optical lattice potentials, Phys. Rev. Lett. 91, 010407 (2003). They are also available upon reasonable request. [37] M. Anderlini, J. Sebby-Strabley, J. Kruse, J. V. Porto, W. D. Phillips, Controlled atom dynamics in a double- well optical lattice, J. Phys. B 39, S199 (2006). [38] R. G. Parr, W. Yang, Density-Functional Theory of Atoms and Molecules (Oxford University Press, New York, 1989). [39] A. Omran, et al., Microscopic observation of pauli block- ing in degenerate fermionic lattice gases, Phys. Rev. Lett. 115, 263001 (2015). [40] V. P. Gupta, Principles and applications of quantum chemistry (Academic Press, 2015). [41] M. H. Schleier-Smith, I. D. Leroux, H. Zhang, M. A. 10

On-site Attractive Fermionic Raman Stage interaction potential hopping coupling I(a) U → 0 0 0 0 I(b) 0 V0 0 0 I(c) 0 → UV0 0 0 II UV0 0 → tF 0 III UV0 tF 0 → g

Extended Data Table 1 | Evolution of the main pa- rameters of the system during the adiabatic prepa- ration based on steps I-III presented in the Methods. To simplify the preparation of |ψ1i, the step I illustrated in the Extended Data Fig. 2 has been divided in three consecutive substeps: I(a). Starting from state |ψ0i, the on-site interac- tion, U → 0 is adiabatically canceled. This brings the Mott excitation into a delocalized single excitation shared by all the P † 3/2 atoms in the insulator, j aj |0i/N . I(b). As proposed in the Methods, one can use a dynamic double-well potential to move the fermionic wavepackets in opposite directions. Note that, as fermionic and bosonic species are decoupled at this point (U = 0), the Mott excitation will remain in the symmet- ric state reached in I(a). I(c). We adiabatically restore the on-site repulsion U → 0. The bosonic state then evolves to a superposition localized at the new position√ of the fermionic † † atoms, (an1 + an2 )|0i/ 2. 11

a b

ndiv = 3

N

j (sites)

Extended Data Fig. 1 | Results of the G-S algorithm. We apply the G-S algorithm to identify the phase mask associated to a holographic 3D Coulomb potential on a lattice of N 3 sites. Fixing the origin at the central site, we choose the nuclear −1 position as n = [(2nndiv) , 0, 0] (note the shift on the first coordinate, so that the lattice induces a natural cutoff). In a we plot an axial central cut on direction z (see aligned set of red sites in b) of the the potential created by a phase mask composed of (ndivN) × (ndivN) cells for ndiv = 3 (see inset). In step (ii) of the algorithm, the Ewald sphere is discretized using a parallel projection, as used in Ref. [32]. The field is inititiated with random phases. Parameters: N=30 and 7000 iterations of the G-S algorithm.

Mott fermionic excitation density

a b I(b) I

I(a) I(c) II Hqc H H III qc e-e

Extended Data Fig. 2 | Numerical simulation of the adiabatic preparation of the ground state of Hsim associated to H2 (particularized for a bidimensional lattice). Red dashed lines follow the adiabatic evolution of the initial state |ψ0i, and arrows point to the evolution direction. a, Shows the preparation of the bosonic state through steps I(a)-I(c) (see Methods and Extended Data Table. 1). Here, continuous lines indicate the exact energy of the two lowest energy sates. For the adiabatic evolution we have Trotterized time as ∆t · U = 0.5 and evolved with |∆U| /(U 2 · ∆t) = 3 · 10−4. b, Continues the preparation of the fermionic state following steps (II-III). In the simulation, we have Trotterized the time evolution in intervals of ∆t·V0 = 0.05. 2 In (II), the kinetic term is adiabatically introduced in steps of ∆tF /(V0 · ∆t) = 0.005. In (III), the electronic repulsion is tuned 2 up as ∆V/(V0 · ∆t) = 0.02. Here yellow (blue) continuous lines follow the exact energy levels of Hqc calculated by imaginary time evolution with(out) the effect of electronic repulsion. Panels show the Mott excitation/fermionic population in the lattice at the key moments indicated in the figure. The final point of the evolution shown in b corresponds to d/a0 = 1. Parameters: N = 60, U/Jc = 1.5, d/a = 10. 12

SUPPLEMENTARY MATERIAL the main text for a single electron. The first limit stems from the fact that more than one lattice site is required Here we provide details about the results and scalings to properly describe an atomic orbital. The upper-bound stated in the main text. It is structured as follows. In of the inequality guarantees that the orbitals fit into the SectionI, we describe the main effect of solving the quan- simulator volume to prevent finite-size effects, where we 1/3 tum chemistry problem in a discrete lattice rather than have assumed that the orbital size scales as Ne , based the continuum. We both explain the origin of the in- on the radial electronic density of alkali atoms. Within equality (5), and give the details on how to plot Fig. 2 of this range of parameters, deviations from the analytic the main text. In SectionII we explain how the effective value come from the discretization of the lattice. For in- Coulomb repulsion between fermions emerge in our sim- stance, one can analytically compare the integral in the ulator, and discuss the origin of the inequalities (8-11) continuum, drψ(r)2V (r) to the Riemann sum that we presented in the main text. We finally devote SectionIII are effectively computing for Slater wavefunctions with R r/a0 3/2 to discuss both how to optimally choose the simulator pa- radial density 2e− /(a0) . One obtains that the de- rameters to accurately obtain the H2 molecular potential viation in the energy decreases with the effective Bohr 2 of Fig. 3, and the numerical method employed to obtain radius as (tF /V0)− , which we numerically confirm for the data in the figure. the studied orbitals as illustrated in the inset of Fig. 2a. We observe that deviations smaller than 0.3% can be reached for a lattice size N = 100 and tF /V0 < 5. I. DISCRETIZATION AND FINITE SIZE EFFECTS IN ATOMIC HAMILTONIAN

In this Section we focus on the effects emerging from both discretization and finite size effects. Since the ef- II. EFFECTIVE FERMIONIC REPULSION fect of the fermionic repulsion is discussed in detail in sectionsII andIII, here we restrict to the single particle Obtaining Coloumb repulsion between fermionic atoms Hamiltonian, is the crucial ingredient of our proposal. In this sec- tion, we give full details on how this effective potential Hkin + Hnuc = tF f †fj ZnV ( j rn ) f †fj . emerges, and the conditions (8-11), given in the main − i − k − k j i, j n, j text, that need to be satisfied to derive it. hXi X (S1) The mapping of the typical length/energy scales of the atomic problem, that are the Bohr radius (a0) / Ryd- berg energy (Ry) is given in Eq. 4 of the main text. In Fig. S1 we present a more systematic benchmarking of A. Complete simulator Hamiltonian and analysis these identities, focusing on 1s orbitals of atomic Hy- procedure drogen. Since this is a quadratic Hamiltonian, we can use exact diagonalization to obtain the lowest part of the As described in the main text, the complete simula- spectrum of a single electron and a single nucleus, i.e., tor Hamiltonian is composed by three terms: H = an Hydrogen atom. This is also what we do in Fig. 2 sim Hkin + Hnuc + HM , namely, the kinetic energy term cor- of the main text for several ratios of tF /V0, i.e., for responding to the hopping of fermionic atoms to nearest several expected atomic sizes. To avoid the divergence neighbour sites with rate J , the nuclear potential at- of the nuclear potential, we chose its central point as1 F traction of Hnuc, and HM , given in Eq. 5 of the main (m, m + 1/2, m), which induces a natural cut-off of the text, which contains both the Mott insulator dynamics divergence. Here m = N /2 , where ( ) represents b M c b·c d·e and the interaction between the fermionic and Mott in- the floor (ceiling) function. We also use open boundary sulator atoms. conditions in all the figures of the manuscript. To appro- priately compare the energies with the continuum limit, The way we analyze the problem is to use a Born- Oppenheimer approximation for the fermionic atoms, we shift the extracted energies by 6tF and finally divide that is, assuming their timescales to be much slower by Ry to express the result in atomic units. As the Bohr radius increases, we observe that the nu- than the Mott insulator atoms dynamics. Like this, merical result approaches the analytical value in the re- we can first solve the Mott-insulator Hamiltonian for 1/3 each fermionic configuration, and then, self-consistently gion 1 < a /a < N/N , that is, the (5) inequality of 0 e find both the effect of the Mott-insulator in the fermion dynamics and the conditions under which this Born- Oppenheimer approximation is valid. 1 By (·, ·, ·) we denote coordinates in the lattice written in units Thus, for a given configuration of the fermionic atoms, of the lattice spacing, a. Integer coordinates correspond to the denoted as, j = j1,..., jNe , we make the follow- lattice sites. ing decomposition{ } of the Mott-insulator Hamiltonian 13

a b hand, it contains two degenerate subspaces with 3 N N ∗ := N N a -states of energy ∆+U [∆] that e M M − e | i we denote as U [ U ], respectively. On the other hand, the subspaceA A\Aformed by the b -excitations can be diagonalized if we imposeB periodic| boundaryi conditions 3 ik j and define the operators bk† = (1/ NM ) j e− · bj†, which gives rise to the energy dispersion: ωk = p P 2J(cos(kx) + cos(ky) + cos(kz)). Now, we treat H1 as a perturbation of H0, focusing on c d the U subspace which is the one coupled directly to the fermions.A Since this is a degenerate subspace, we must /a 0

a apply degenerate perturbation theory [42]. This theory assumes that the perturbation, H1 in our case, breaks the degeneracy in H0, and starts by choosing an appropriate basis compatible with these new eigenstates. In our case, it is convenient to choose a basis for privileging the symmetric combinations of excitationsA in the and AU U subspaces, that can be generally written as: FIG. 1: Validity of atomic units in finite systems. b, A\A Radial density of sates associated the ground state of Hqc (0) 1 ϕ = a† 0 . (S4) in a cubic lattice with N = 80 sites per side for different X j | i dim(X) j X values of tF /V0 (in a we show an axial cut for tF /V0 = 5). E X∈ After normalizing by the central density, log scale reveals the p −r/a0 where X is the subspace or , and dim(X) exponential radial decay e characteristic of 1s orbitals AU A\AU before finite-size effects appear. A linear fit of the first 30 the dimension of this subspace. In particular, we are values provide the decay parameter 1/a0. In c, we plot this interested in the corrections to the symmetric state in extracted Bohr radius for different sizes of the lattice and (0) U , ϕ U , which is the one that mediates the correct compare it to the scaling in the continuum, a0/a = 2tF /V0 A A repulsive potentialE between the fermions. Thus, we need (dashed line). The critical Bohr radius at which finite-size effects conspire with the scaling increase linearly with N. This to calculate: effect is also appreciated in d when we compare the energy of this ground state to the dependence in the continuum, Ry = 2 (0) (1) (2) V0 /(4tF ) (dashed line). Es, U = Es, U + Es, U + Es, U + ..., A A A A (0) (1) (2) ϕ U = ϕ U + ϕ U + ϕ U + .... (S5) | A i A A A E E E HM ( j ) = H0 + H1, where each term reads as, { } As we will show afterwards, Es, U will translate into an effective potential between theA fermions, while ε = H0 = U a†aj + ∆ a†aj + J b†bj , (S2) j j i 1 ϕ(0) ϕ 2 will be a measure on how much we j j i,j U U X{ } X Xh i deviate− |h A from| A thei| ideal dynamics. Thus, we will impose 3 H1 = Jc/N a†aj + g a†bj + b†aj . (S3) ε 1 to derive the conditions (8-11) on the main text. M i j j  Xi, j Xj   Here, we will explicitly write the inverse dependence of B. Breaking the symmetric state degeneracy with 3 the cavity coupling with its volume (NM ). cavity coupling Notice that H ( j ) conserves the total number of M { } excitations, j aj†aj + bj†bj . Thus, if we initialize our As prescribed by degenerate perturbation theory, we simulator with a single excitation in Mott-insulator, the calculate the perturbed energies/wavefunctions in two P   dynamics will be restricted to the single-excitation sub- steps by separating the contribution of the cavity Hamil- 3 tonian H1,c = Jc/NM i,j ai†aj, from the rest of the space formed by / = aj†/bj† 0 j, where 0 is the A B { | i} | i perturbation H1,g = H1 H1,c. The cavity Hamilto- state with no excitations in the / atomic states. The P− intuition is that the the on-site potentialA B provided by the nian is enough to break the degeneracy of the symmetric fermions at positions j localizes the Mott-insulator ex- states of U [ U ], giving them an extra energy ρM Jc { } [(1 ρ )AJ ], respectively.A\A Here, we define ρ N /N 3 . citations around them forming a bound state. As we will − M c M ≡ e M show, the energy of this bound state will depend on the However, apart from breaking the degeneracy it also cou- particular fermionic configuration, j , in the same way ples both symmetric states with strength: than the Coulomb potential appearing{ } in chemistry. ϕ(0) H ϕ(0) = ρ (1 ρ )J √ρ J , To show it, let us first analyze the structure of U 1,c U M M c M c A\A A − ≈ H0 in the single-excitation subspace: on the one D E (S6) p

14

3 2 where in the last approximation we use Ne NM . Thus, 6J, which requires the following inequality: Ne V0 in order to keep the mixing between symmetric states JN(a/L)2. In this estimation, we are using a simpli- small, we must impose U Jc, that is inequality (8) of fied formula for electron-electron repulsion [43], Ve e  − ≈ the main text. This guarantees that: V (N 1)2/3 ρ4/3(j) , and approximate an homoge- 0 e − j (1,c) neous electron density, ρ(j) ρ, for the optimal situation Es, U ρM Jc , (S7) P ≈ A ≈ in which molecular orbitals fully occupy the lattice. This and that the first correction to the eigenstate is: gives the (10) inequality given in the main text. Finally, let us estimate how much the wavefunction (1,c) Jc√ρM (1,c) (0) ϕ ϕ . (S8) ϕ gets perturbed by H1,g. To first order: U U s, U A ≈ U A\A A E E E Notice, that since ρ 1, the inequality (8) of the (1,g) M ϕ = β b† 0 , (S12)  s, U k k main text already guarantees that Jc√ρM /U 1 and A | i  E k (c) X bounds higher order correction in Es, U . A 2 whose norm is given by, k βk , which can be explicitly computed as, | | P C. Non-degenerate perturbation theory with H1,g

2 2 U 2 ikj1 ikjNe 0 bkH1,g ϕA g 1 e + ... + e Once H1,c breaks the degeneracy, we can apply non- s h | (0) | i = 3 2 . degenerate perturbation theory for the rest of the per- Ne NM (U + ∆ + ρM Jc ωk) k Es U ωk k X A − X − turbation H1,g. To second order, we find: (S13)

2 This sum can be calculated taking energy derivatives in 2 ikj1 ikjNe (2,g) g 1 e + ... + e (S10), observing that the resulting condition is already E = . (S9) s, U 3 guaranteed by imposing the most restrictive condition A Ne NM U + ∆ + ρM Jc ω k Xk − (10), that we derived by calculating the exact energy of In the continuum limit, that is, when N 1, we can the bound state. M  transform the sum into an integral and find: Second order perturbation in H1,g for the wavefunction 2 ikr leads to two type of contributions: g 1 e V0 r/L e− , 3 (2,g) s⊥ s⊥ Ne NM U + ∆ + ρM Jc ωk ≈ 2r/a ⊥ ⊥ k ϕs, = δs U ϕ U + ηs U ϕ , − U A A U X A ⊥ A ⊥ A\A g2 1 aN V E s U E s U E 0.25g2/J e 0 . XA XA\A 3 (S14) NM U + ∆ + ρM Jc ωk ≈ − 2L Xk − namely, the ones that couples the symmetric state to (S10) the antisymmetric ones in the U and U sub- spaces passing by b-states. TheA norm ofA\A the former, where L/a = J/(U + ∆ + ρM Jc 6J) is the effective 2 ⊥ δ ⊥ , can be approximately upper bounded − s U s U length of the Yukawa-type interaction that appears in A | A | p by: (2) P Es, U . This length can be controlled by the effective A 2 detuning between the symmetric state energy U + ∆ + s⊥ s 2 † ϕ U H1bk 0 0 bkH1 ϕ U Ne V0 ρM Jc, and the upper band-edge of ωk at 6J. Notice, h A | | ih | | A i < . 2 (0) (0) (0) 2 ρM Jc that we define V0 = g /(2πNeJ) as the effective repulsion ⊥ E ω E E ⊥ ∼ s k s, U k s, U s , U   X X A − A − A strength.     Before proceeding to compute the correction to the Thus, by imposing it is small we obtain one of the wavefunction, let us compare the result of the per- conditions (9) of the main text, i.e., V ρ J /√N . 0  M c e turbative expansion of the energy with the exact so- The contribution of the antisymmetric state of U , 2 A\A lution of Es, U . The latter can be obtained in our ⊥ that is, s⊥ ηs U is already small provided that A U | A | problem by directly solving the Schr¨odinger equation, the previousA\A inequalities are satisfied, such that it does (H0 + H1) Ψs = Es Ψs just imposing that Ψs lies not introduceP any new condition. in the single-excitation| i | subspacei of + . In| thei con- ditions of the text, the result is givenA inB the following closed equation: D. Effects on the fermionic dynamics 2 g2 1 eikj1 + ... + eikjNe Once we have calculated the energy of the bound state Es, U U+∆+ρM Jc+ 3 . A ≈ Ne N Es, U ωk M k A − for each fermionic configuration, j , we now have to X (S11) project the fermionic Hamiltonian:{ } By comparing both expression, we see that the sec- ond order correction introduced by coupling to the b- H = J f †f + H + U a†a f †f , (S15) f − F i j nuc j j j j atomic states must be smaller that U + ∆ + ρM Jc i, j j − hXi X{ } 15 into the basis that combines the N fermions, with the e a b symmetric spin excitation which mediates the atomic po- tential, that is:

s j ,... j = f † . . . f † 0 ϕ , (S16) critical 1 Ne j1 jN j1...jNe || ii e | i ⊗ Bohr radius   E for j1 ,... jNe any configuration for the position of the Ne fermions. To lowest order we obtain,

˜ Hf tF fi†fj + Hnuc + He e , (S17) ≈ − − FIG. 2: Calculation of the critical Bohr radius for H2. i, j hXi a, For each internuclear separation d/a0, we benchmark sep- arations on the lattice d/a ranging from 1 to 30 sites, tuning where one gets both the electron repulsion Hamiltonian, the Bohr radius, a0/a, accordingly. In the figure, d = 1.4a0, with a Yukawa potential He e, and a kinetic energy term close to the equilibrium distance. Similarly to what happened − with a reduction of the tunneling rate, JF , given by the in the atomic case (see Figure 2a), the calculated potential Franck-Condon overlap: polynomially approaches the exact solution (dashed line) as the effective Bohr radius increases, up to the point in which s s Ne 1 finite size effects appear. To calculate this optimal tF /V0, tF = JF ϕ ϕ JF − . (S18) j1+1...jNe j1...jNe we repeat the calculation for a bigger system, and detect the h | i ≈ Ne point where both curves depart. b, To identify this point, we fit the energy of the largest lattice to a universal scaling To self-consistently check the validity of the Born- −2 m (tF /V0) + n, (dashed line). For N = 75 and a nuclear Oppenheimer approach, one must certify that Hf does not efficiently populate states which are not in the ba- separation of 1.4a0, the critical point (indicated by the arrow) corresponds to tF /V0 ≈ 3.5, providing an expected precision sis (S16) we use to project, since Hf can take us out of 10−2Ry in the energy of the minimum potential at this from the symmetric spin excitation, To first order in Hf , given distance. the perturbation to our basis states j1 ,... jNe has the form || ii

(1) s⊥ s⊥ A. Choice of simulator parameters j1 ,... jN = φ j1 ,... jN ϕ , || e ii j1...jNe | e i ⊗ j1...jNe s⊥ E X (S19) In Figure 3a we calculate the energy of the bound state

We can upper bound the perturbation by summing up of the spin excitation as a function on the interfermionic over the different configurations to arrive to: separation. We use exact diagonalization of HM ( j ) for two fermions at positions (m r/(2a) , m, m){ and} 2 − d e ⊥ J (m + r/(2a) , m, m). To calculate the constant energy φs 2 < N F , (S20) b c j1...jNe nn shift C, we also obtain the energy of the spin excita- | | ρM Jc j s⊥ ∼   X{ } X tion bound state when a single fermion is placed at po- sition (m, m, m), and subtract it from the energy of the where Nnn denote the number of nearest neighbors in the two fermions. In this Figure, we choose the parameters lattice (6 in the case of a simple cubic geometry). Im- such that the inequalities (5,8-11) are satisfied and obtain posing that this correction is small, we obtain the other Coulomb repulsion between fermions for NM = 200. inequality (9) of the main text, that is, J ρ J √N . F  M c e To plot the molecular potential of Fig. 3b, that is, Finally, there are two additional inequalities related to to calculate the electronic energy, E(d) as a function of obtaining a truly Coulomb repulsion rather than Yukawa. the internuclear distance d, we center the nuclear poten- On the one hand, the length of the bound state has to tial in positions r1 = (m d/(2a) , m + 1/2, m) and ideally be larger than the fermionic lattice, L aN, − d e  r2 = (m + d/(2a) , m + 1/2, m) and obtain the ground while being smaller than the Mott insulator lattice, L < state energyb usingc the numerical methods explained in aNM , such that the bound state does not feel the border. the next Section. Since two electrons are involved, the Combining both inequalities, we arrive to the condition extracted energy is now shifted by 12tF , and finally writ- (11) of the main text. ten in atomic units. Notice also that, since we use spinless fermions, we have to restrict to the symmetric subspace of the electronic problem so that we can compare the re- III. OBTAINING MOLECULAR POTENTIALS sults to those of the H2 molecule, which is formed for two spins of opposite sign. In this section, we first give more details on how we As it happens in the atomic case, accuracy increases choose the simulator parameters to plot the figures of as the Bohr radius grows up to a critical value at which the manuscript, and explain the numerical methods em- finite-size effects are relevant. However, the optimal ployed for the two-electron wavefunction calculation. choice of the Bohr radius now depends on the number 16 of lattice sites that separate the nuclei. To identify this Once we have built an approximated basis, we need critical Bohr radius at which the finiteness of the lat- to project the Hamiltonian Hqc into the basis. The tice N = 75 compromises the accuracy, we use the fol- terms associated to the kinetic energy and nuclear inter- lowing procedure. First, for a given internuclear sepa- actions are easily projected, as they only depend on single ration d/a we solve the electronic structure for nuclear fermionic orbitals, φi φj H0 φr φs = δjs φi H0 φr + 0 h | | i h | | i potentials separated a number of lattice sites d/a rang- δir φj H0 φs . The lattice imposes a natural cutoff h | | i ing between 1 and 30, increasing the Bohr radius accord- (S10), corresponding to V (0) πV0, and the main diffi- ≈ ingly, tF /V0 = (d/a)/(2d/a0). The same calculation is culty comes from calculating terms associated to e-e in- repeated for a bigger system size, N = 100. Both lat- teractions, Hee. At a first glimpse, they involve a sum of tices provide compatible results as long as finite-size ef- N 6 coordinates, fects are not important. The point where both curves deviate corresponds to an approximate optimal t /V , V (r r )φ (r )φ (r )φ (r )φ (r ) , (S23) F 0 1 − 2 i 1 r 1 j 2 s 2 that provides maximum accuracy for the lattice size con- r1,r2 X sidered, e.g., N = 75 in our case. In practice, we choose where, V (r) = V / r . In the reciprocal space, however, the point at which finite-size energy deviations are one 0 k k order of magnitude smaller than the discretization error this sum simplifies as (see Fig. S2). Fitting these values for each internuclear separation d/a, we choose t /V = 4.2 d/a 0.065, i.e., as φi φj Hee φr φs = V (k) (φ^i φr)(k) (φ^j φs)( k) , F 0 − · h | | i · · · · − the nuclei are more separated, the border of the system k X linearly approaches, needing to reduce the Bohr radius e (S24) 3 accordingly. and only N terms are involved, speeding-up the calcu- lation. Here f denotes the Fourier transform of function f. In principle, this induces periodic boundary conditions in the system,e which are undesirable as fermions would B. Numerical methods to deal with the interact along the minimum distance measured on the two-electron problem periodic lattice, overestimating e-e interactions. To solve this issue, we double the size of the system before calcu- lating the Fourier transform, and impose null probability Another important difference with respect to the single of occupying these artificial positions. Fourier transforms electron problem is that exact diagonalization strategy is are obtained using a Discrete Fast Fourier Algorithm. out of reach for our computational resources. To prevent This last calculation is the bottleneck from the compu- this situation, we artificially reduce the degrees of free- tational time perspective and, at the expense of memory, dom by projecting the Hamiltonian on a single fermion we initially store the FFT for each of the n(n + 1)/2 basis, φ n . This projected Hamiltonian reads as, { i}i=1 product of pairs of molecular orbitals, so that the trans- formation is not unnecessarily repeated. It is also useful n to note that not every term hijrs needs to be calculated, H = h φ φ φ φ , (S21) P ijrs | i ji h r s| due to the symmetries of the Hamiltonian. For example, i,j,r,s=1 X h1123 = h1132 = h2311 = h3211. In practice, this reduces the calculated terms from n4 to n2(n2 + 3)/4 indepen- where φi φj denotes the product state φi φj and dent terms, where n is the number of molecular orbitals h =| φ φi H φ φ . | i ⊗ | i in the projected basis. For Figure 3b, we observed that ijrs h i j| eff | r si The success of this strategy now depends on how accu- convergence in energies to the order of the discretiza- tion effects was reached for 15 Hartree-Fock orbitals and rately the orbitals in this set can describe the interactions + in the Hamiltonian. We then choose a basis composed by 15 low-energy H2 states. This corresponds to n = 30, + 203175 independent terms, and approximately 8h of total single-fermion states of H2 calculated with exact diago- nalization, together with more orbitals obtained using a computational time in a 2.20GHz CPU. Hartree-Fock approximation [22]. This is, starting with The mean charge interaction in the Hartree-Fock cal- the ground state obtained for a single fermion, we iterate culation can also been rewritten as, the equation, 2 j i 1 2 φ V ( i j ) φ = φ − V (k) φ ( k) , 12 k − k · h |F · − i 2 j ( φ )i + φj V ( j i ) φi = λ φi , (S22) X    H0 | i 12 k − k e f (S25) j 1 X  where − denotes the inverse Fourier transform. We shouldF emphasize that this projection on a single-particle until convergence is reached. This way, orbitals inter- basis is just a numerical strategy that enables us to nu- act with a mean charge induced by the rest of fermionic merically benchmark the model, but does not have any atoms in the lattice, while we neglect the exchange inter- impact on the experimental implementation of the pro- action. posed analog simulator. 17

[42] J. J. Sakurai, J. Napolitano, Modern quantum mechanics (Addison-Wesley, 2011). [43] R. G. Parr, Density-Functional Theory of atoms and molecules (Horizons of Quantum Chemistry. Springer, Dordrecht, 1980).