The strong spin-orbit interaction in the organic-inorganic perovskites tied to the incorporation of heavy elements (e.g., Pb and I) makes these materials interesting for applications in spintronics. In conjunction with a lack of inversion symmetry associated with distortions of the metal-halide octahedra, surfaces and interfaces, or the application of a bias, the Rashba effect (used in spin field-effect transistors and spin filters) has been predicted to be much larger in these materials than in traditional III-V semiconductors such as GaAs. Evidence of strong Rashba coupling has been observed in both 3D (bulk) and 2D perovskites, with the relative role of bulk and surface Rashba contributions in the former case under active debate. The varying size of the reported spin splittings points to the need for more experimental studies of Rashba effects in the organic-inorganic perovskite family of materials. Here, we apply time-resolved circular dichroism techniques to the study of carrier spin dynamics in a 2D perovskite thin film [(BA)2MAPb2I7; BA = CH3(CH2)3NH3, MA = CH3NH3]. Our findings confirm the presence of a Rashba spin splitting via the dominance of precessional spin relaxation induced by the Rashba effective magnetic field (also known as D’yakonov Perel spin relaxation). The size of the Rashba spin splitting in our system was extracted from simulations of the measured spin dynamics incorporating LO-phonon and electron-electron scattering, yielding a value of 10 meV at an electron energy of 50 meV above the band gap, representing a 20 times larger value than in GaAs quantum wells.

Hybrid organic-inorganic perovskites have gained considerable attention in recent years due to their outstanding performance as absorbing layers in photovoltaics.1 This success has led to a comprehensive research effort with the aim to unravel their photophysical properties2–18 and to move beyond CH3NH3PbI3 and explore other material compositions including 2D perovskites.19–24 This burgeoning family of materials offers properties that can be tailored to a broad range of applications in optoelectronics, including photovoltaics,1,23,25 field-effect transistors,26,27 hard radiation detectors,28 light-emitting diodes,29–31 lasers,32 and optical sensors.33 

Hybrid perovskites are characterized by strong spin-orbit coupling (SOC) tied to the constituent heavy elements.34 These strong spin-orbit effects make the perovskite family of materials attractive for applications in semiconductor spintronics and spin opto-electronics.35–45 SOC leads to a giant (∼1 eV) splitting of the lowest two conduction bands and influences the band gap and carrier effective masses.46,47 In conjunction with a lack of inversion symmetry, SOC also leads to an effective magnetic field that lifts the degeneracy of the carrier spin states within each band.35 While this effect has many origins in semiconductors tied to different sources of inversion asymmetry,48–52 it is most commonly referred to as the Rashba effect after Bychkov and Rashba analyzed the case of structural inversion asymmetry in a two-dimensional electron gas.50 In a spintronic device, the effective magnetic field associated with the Rashba effect may be used to operate on the spin state of carriers by inducing precession in a spin field-effect transistor.36,37 The Rashba spin splitting can also enable spin-dependent transport without the need for an external magnetic field or magnetic materials by using a resonant tunnel diode structure.53 

Rashba effects in organic-inorganic perovskites have been the subject of numerous theoretical investigations in recent years.54–60 The predicted Rashba splittings (∼10’s of meV) are large compared to the conventional III-V semiconductors upon which most spintronic devices have been based.34 Switchable Rashba coupling exploiting the ferroelectric response has also been predicted.56,57 Experimental observation of the Rashba effect in the valence band in single crystal CH3NH3PbBr3 was recently demonstrated using angle-resolved photoemission spectroscopy (ARPES).61 The magnitude of the observed Rashba spin splitting (160 meV) was comparable to the largest values predicted theoretically in any of the hybrid perovskite systems.54–60 While it has been suggested that this observed Rashba effect is due to the symmetry breaking at the surface,62–64 since the strong dipole tied to surface reconstruction is expected to penetrate several hundred nanometers,65 the ultralarge splitting observed is, nevertheless, promising for developing thin-film based spintronic devices. Experimental evidence of Rashba effects have been seen in perovskites using the photogalvanic effect,66 transient anisotropy,67 and electro-absorption68 in recent years, highlighting the need for further experimental studies of Rashba effects in the hybrid perovskite family of materials.

Here, we report detection of the Rashba splitting in the 2D perovskite (BA)2MAPb2I7 through measurement of carrier spin relaxation using time-resolved circular dichroism techniques. Measurements of carrier spin dynamics have proven extremely valuable in the study of Rashba coupling in inorganic III-V semiconductor systems,35 in which the spin population relaxation time (T1) is sensitive to both the magnitude and wavevector-dependence of the Rashba effective magnetic field.48,52 (BA)2MAPb2I7 exhibits a Ruddlesden-Popper structure with two layers of corner-sharing lead-iodide octahedra separated by long organic cations of butylammonium forming an interdigitating bilayer. Our experiments indicate that T1 decreases with increasing photoexcitation energy and increases with optically injected carrier density, both signatures of Rashba precessional spin relaxation.48,69,70 Our simulations of the electron spin dynamics yield a Rashba spin splitting of 10 meV at an electron energy 50 meV above the band gap, a value that is 20 times larger than in GaAs quantum wells.48 Our findings highlight the utility of spin relaxation studies as a probe for Rashba coupling in the organic-inorganic hybrid perovskite family of materials.

In systems with no center of inversion symmetry, the spin-orbit interaction leads to an effective magnetic field Ωk with a direction that determines an effective spin quantization axis at each value of the electron wavevector k and a magnitude that determines the energy separation between spin-up and spin-down band states. The crystal structure will dictate the wavevector dependence of Ω(k). For the 2D organic-inorganic perovskites, symmetry breaking is tied to tilting and associated polar distortions of the metal-halide octahedra, which are determined by the interplay between the long organic BA cations and the smaller MA cations that induce out-of-plane and in-plane octahedral distortions, respectively.24 For the n = 2 structure studied in this work [Fig. 1(a)], x-ray diffraction indicates a noncentrosymmetric structure with a polar (C2v) space group at room temperature.24,71 As a result, a nonzero Rashba spin splitting is expected for this system. The Rashba Hamiltonian is given by50,58

(1)

where k = (kx, ky) is the electron wavevector in the x-y plane perpendicular to the stacking direction, σ is the vector of Pauli spin matrices, and

In this case,

(2)

and the spin eigenstates are given by

(3)

where λR is the Rashba coupling parameter. The Rashba effective magnetic field and the dispersion relations for the two spin eigenstates are shown in Fig. 1(b).

FIG. 1.

(a) Linear absorption of the sample studied. Upper inset: (BA)2MAPb2I7 crystal structure; lower inset: Photo of thin film sample on sapphire substrate. (b) Upper panel: Rashba effective magnetic field Ω(k) showing the variation of the direction at a fixed value of |k|. Lower panel: Energies of the spin eigenstates (E+: red curve, E: blue curve) as a function of in-plane wave vector. (c) Optical selections rules for interband transitions at the band gap of (BA)2MAPb2I7.

FIG. 1.

(a) Linear absorption of the sample studied. Upper inset: (BA)2MAPb2I7 crystal structure; lower inset: Photo of thin film sample on sapphire substrate. (b) Upper panel: Rashba effective magnetic field Ω(k) showing the variation of the direction at a fixed value of |k|. Lower panel: Energies of the spin eigenstates (E+: red curve, E: blue curve) as a function of in-plane wave vector. (c) Optical selections rules for interband transitions at the band gap of (BA)2MAPb2I7.

Close modal

The Rashba effective magnetic field leads to precessional carrier spin relaxation. This mechanism was first discussed in the context of III-V semiconductors by D’yakonov and Perel69 and later refined for semiconductor quantum wells by D’yakonov and Kachorovskii.70 Due to the optical selection rules [Fig. 1(c)], excitation with circularly polarized light leads to a fully spin polarized distribution of electrons and holes. The carrier spins will be initially aligned parallel or antiparallel to the light propagation direction [the z direction in Fig. 1(b)] depending on the carrier type (electron or hole) and the circular polarization state of the light field. These initially polarized carrier spins will begin to precess about Ω(k) with frequency |Ω(k)|. Since the direction of Ω(k) varies with k, the net spin polarization of the ensemble, after summing over k decays. Our interest here is on the longitudinal spin lifetime in the absence of an external magnetic field (T1). The spin relaxation rate is

(4)

where τ is the carrier scattering lifetime. A characteristic feature of this spin relaxation mechanism is that T1 is proportional to the carrier scattering rate,70 referred to as motional narrowing. The instantaneous magnitude and direction of Ω(k) change as carriers scatter from one k value to another. The larger the rate of scattering, the slower the spin relaxation because the carrier spins do not have as much time to undergo precession between scattering events. This mechanism of spin relaxation is dominant in III-V semiconductors such as GaAs at room temperature.35,48

Since the rate of precessional spin relaxation is proportional to the square of the Rashba effective magnetic field,48 measurement of T1 provides a sensitive probe of the strength of Rashba coupling in a semiconductor material. For instance, in GaAs quantum wells, the spin lifetime at 300 K is 100 ps72 and the corresponding spin splitting has been calculated to be <1 meV.48 In contrast, in quantum wells formed from the InAs/GaSb/AlSb system possessing a much larger SOC, calculated spin splittings are ∼10 meV or larger,48,53 and measured spin lifetimes in these materials under 1 ps have been reported.42,52

The results of differential transmission measurements [(TT0T0), where T (T0) is the transmission of the probe pulse in the presence (absence) of the pump beam, see Fig. 2(a)] for linearly polarized pump and probe pulses are shown in Fig. 2(b). The use of linearly polarized pulses enables measurement of the average state filling signal associated with both spin populations. The magnitude of the differential transmission signal at τ = 2 ps is plotted alongside the linear absorption spectrum in the lower panel of Fig. 2(b). A Tauc analysis of the linear absorption results (see the supplementary material) indicates a band gap of (2.12 ± 0.05) eV. The peak at 2.05 eV is therefore attributed to the state filling response of excitons, and the peak at higher energies is due to unbound electron-hole pairs on interband transitions above the band gap. Fitting the results in Fig. 2(b) for the above band gap excitation to a double exponential decay yields a carrier thermalization time of (7 ± 1) ps, followed by recombination on a time scale ∼1 ns. These spin-independent relaxation kinetics are in line with previous studies in phenylethylammonium lead iodide (PEA)2(MA)n−1PbnI3n+1 2D perovskite.73 

FIG. 2.

(a) Schematic diagram of the time-resolved circular dichroism experiment. (b) Upper panel: Differential transmission (TT0T0) as a function of laser tuning and interpulse delay with orthogonal linear polarizations in the pump and probe pulses to probe the spin-independent carrier dynamics. Lower panel: The magnitude of the TT0T0 signal at 2 ps delay, plotted together with the linear absorption spectrum. For these results, the pump pulse fluence was 0.12 μJ/cm2. The differential transmission signal indicates separate peaks tied to excitons (2.05 eV) and unbound electron-hole pairs above the band gap (2.18 eV). The sharp distinction between exciton and interband responses in the nonlinear optical signal reflects the fact that the bleaching response is not sensitive to scattering losses, unlike the linear absorption response. (c) Results of circular dichroism experiments for a laser tuning of 2.15 eV. The blue (red) curve shows the results for the same (opposite) circular polarization states in the pump and probe beams. Inset: Degree of spin polarization vs delay. The red curve indicates the results of fitting to a single exponential with a decay time of (10 ± 1) ps.

FIG. 2.

(a) Schematic diagram of the time-resolved circular dichroism experiment. (b) Upper panel: Differential transmission (TT0T0) as a function of laser tuning and interpulse delay with orthogonal linear polarizations in the pump and probe pulses to probe the spin-independent carrier dynamics. Lower panel: The magnitude of the TT0T0 signal at 2 ps delay, plotted together with the linear absorption spectrum. For these results, the pump pulse fluence was 0.12 μJ/cm2. The differential transmission signal indicates separate peaks tied to excitons (2.05 eV) and unbound electron-hole pairs above the band gap (2.18 eV). The sharp distinction between exciton and interband responses in the nonlinear optical signal reflects the fact that the bleaching response is not sensitive to scattering losses, unlike the linear absorption response. (c) Results of circular dichroism experiments for a laser tuning of 2.15 eV. The blue (red) curve shows the results for the same (opposite) circular polarization states in the pump and probe beams. Inset: Degree of spin polarization vs delay. The red curve indicates the results of fitting to a single exponential with a decay time of (10 ± 1) ps.

Close modal

T1 was measured using time-resolved circular dichroism. A circularly polarized pump pulse is used to inject a spin-polarized carrier distribution. When the probe has the same (opposite) circular polarization as the pump, the state filling signal is tied to the majority (minority) spin population. The convergence of these two state filling signals with increasing interpulse delay indicates carrier spin relaxation. Figure 2(c) shows the results of circular dichroism experiments for the same pump pulse fluence as in Fig. 2(b) under excitation at 2.15 eV (30 meV above the band gap). A strong oscillatory signal associated with the coherent artifact74 is superimposed on the state-filling response for delay values within the range of overlap between the pump and probe pulses. Beyond this overlap region, the differential transmission signals for the same circular polarization (SCP) and opposite circular polarization (OCP) are clearly resolved, indicating a difference in the magnitude of the state-filling response tied to the majority and minority carrier spin populations. The degree of spin polarization of the carrier distribution may be extracted from the results in Fig. 2(c) by taking the difference of the SCP and OCP responses and dividing by the sum (see the supplementary material for further details). The result of this analysis is shown in the inset of Fig. 2(c). The decay of the carrier spin polarization fits to a single exponential, yielding T1 = (10 ± 1) ps. Recent calculations indicate similar effective masses for electrons and holes in this system.75 The resulting similarity in the density of states in the valence and conduction bands has the consequence that the two carrier species contribute to a similar extent to the magnitude of the measured state filling signal. The observation of a single exponential decay, together with the large value of the initial degree of spin polarization, points to a similar spin lifetime for electrons and holes.

The measured spin lifetimes for a range of laser tunings (Eex) and pump pulse fluence values are shown in Figs. 3(a) and 3(b), respectively. In order to detect evidence of Rashba precessional spin relaxation, the spin lifetime for carriers above the band gap is of interest. Figure 3(a) shows the spin lifetime vs laser detuning (EexEg), which together with the carrier masses determines the electron and hole energies relative to the band edges. The spin lifetime decreases for increasing laser detuning, indicating that spin relaxation is more rapid for carriers with larger kinetic energies. The spin lifetime was also observed to increase with increasing pump pulse fluence [Fig. 3(b)]. Increasing the density of optically injected carriers therefore results in a longer spin lifetime.

FIG. 3.

Measured dependence of the spin lifetime on laser detuning (a) and laser pulse fluence (b). The solid curves show the results of simulations of the carrier spin dynamics taking the Rashba coupling strength as a fitting parameter. (c) Calculated conduction band spin splitting using the optimum value of the Rashba coupling strength determined from the numerical simulations, indicating a spin splitting of 10 meV at an energy of 50 meV above the band edge.

FIG. 3.

Measured dependence of the spin lifetime on laser detuning (a) and laser pulse fluence (b). The solid curves show the results of simulations of the carrier spin dynamics taking the Rashba coupling strength as a fitting parameter. (c) Calculated conduction band spin splitting using the optimum value of the Rashba coupling strength determined from the numerical simulations, indicating a spin splitting of 10 meV at an energy of 50 meV above the band edge.

Close modal

The fluence and detuning energy dependences in Figs. 3(a) and 3(b) indicate precessional spin relaxation—the D’yakonov Perel mechanism. Since the precession vector Ω(k) has a magnitude and direction that varies with wavevector, precession causes the initially spin-polarized distribution of carriers to decay. Scattering of carriers with phonons, impurities, and each other will slow this spin relaxation process by inducing rapid changes in the value of Ω(k), resulting in a smaller degree of precession between scattering events. This motional narrowing feature is clearly reflected in the observation of an increase in T1 with increasing carrier density, indicating slower spin relaxation with an increased rate of carrier-carrier scattering. The observation of a decrease in T1 with increasing excess energy is also consistent with Rashba precessional spin relaxation. Since the carrier thermalization time was found to be similar to the spin lifetime, the carriers remain hot during spin relaxation, with a temperature determined by the laser detuning. A higher average carrier kinetic energy will lead to the occupation of states at larger k values. Since the magnitude of the effective magnetic field increases with |k|, hot carriers will undergo more rapid spin relaxation.

We note that the results in Fig. 3(b) are opposite to the trend expected for the Elliott Yafet spin relaxation process,76,77 for which scattering events themselves lead to spin decay rather than precession between scattering events. The spin relaxation process tied to spin-flip scattering between electrons and holes (the Bir Aronov Pikus mechanism),78 which would lead to the opposite trend vs excited carrier density to that measured here, is expected to be weak in perovskite materials due to the small value of the exciton exchange splitting,44 a feature that has been attributed to the spatial separation of the electron and hole wave functions within the unit cell.

An estimate of the magnitude of the Rashba spin splitting in the 2D perovskite structure studied here was obtained from simulations of the measured carrier spin dynamics. Calculation of the momentum scattering time enables the strength of the Rashba coupling to be determined from the measured spin lifetime using Eq. (4), where τ is the carrier scattering lifetime associated with all scattering processes neglecting the spin texture. The faster the carrier scattering, the slower the rate of spin relaxation for a given magnitude of the Rashba effective magnetic field. These simulations incorporated polar optical phonon scattering, which is expected to be stronger than acoustic phonon scattering at 300 K,79 as well as 2D electron-electron scattering mediated by the Coulomb interaction in the nondegenerate limit. Electron-phonon scattering and electron-electron scattering both contribute to motional narrowing, decreasing the rate of spin relaxation. In this case, the momentum scattering rate is τ1=τep1+τee1. Since the experiment indicates a similar spin relaxation rate for electrons and holes, only the spin relaxation of electrons in the conduction band was included for simplicity. The polar coupling in 2D can be expressed as

(5)

where α is the dimensionless polar coupling. Note that the polar coupling has different q dependences in 2D and 3D, being 1/q and 1/q, respectively. The momentum scattering rate due to polar optical phonons is

(6)

Here, ω0 is the LO phonon frequency, n0[exp(ħω0/kBT)1]1 is the phonon number, k0=(2meω0/ħ)1/2, Ek is the electron energy, and K(x) and E(x) are the complete elliptical integral of the 1st and 2nd kind, respectively. We emphasize that the apparent divergence of Vq in Eq. (5) at q → 0 does not give rise to any singularity in the scattering rate in Eq. (6), which changes little if a potential Vq that is finite at q = 0 is used.80 

For carrier-carrier scattering, we consider electron-electron scattering via the Coulomb interaction, e2/ϵr, where ϵ is the dielectric constant. For 2D systems, the electron-electron scattering in the nondegenerate limit can be expressed as

(7)

where N is the carrier density and EBmee4/2ϵ2ħ2 is the effective Rydberg energy with me being the electron effective mass. The calculated phonon and electron-electron scattering rates are shown vs electron energy in Fig. S4.

In our calculations, the carrier effective masses me = 0.291m0 and mh = 0.328m0 were taken from recent first principles calculations for (BA)2PbI4,75 the LO phonon energy (ħω0 = 12.4 meV) was taken from Ref. 79, and α = 2.12 was used. The effective dielectric constant ϵ ≡ (ϵala + ϵblb)/(la + lb) is the volume average with ϵa(b) and la(b) being the dielectric constant and width of the inorganic (organic) layer, respectively. Taking ϵa = 6.5 and ϵb = 2 gives an estimate of ϵ = 3 for our structure.3 The detuning energy E is the summation of electron and hole energies. From the momentum conservation, we obtain the electron energy,

(8)

When E < kBT = 25 meV, we take the electron energy to be Ek = kBT/(1 + me/mh). λR was varied to obtain the best fit to the experimental results and was the only adjustable parameter in the simulations. The resulting fits are shown as the solid curves in Figs. 3(a) and 3(b). Quantitative agreement was obtained over the full range of carrier energy above gap and excitation carrier density, yielding λR = 0.08 eV Å. The Rashba spin splitting extracted from the fitting process is shown in Fig. 3(c). A splitting of 10 meV is found at an energy of 50 meV above the conduction band minimum. For comparison, for GaAs quantum wells at 300 K, a much smaller spin splitting of 0.5 meV at this energy was found from the analysis of spin lifetime results.48 The spin-independent pump probe experiments in Fig. 2(b) indicated a thermalization time of 7 ps, resulting in an effective carrier temperature somewhat cooler than determined by the laser detuning alone. Our simulations, which neglect thermalization effects, therefore provide a lower bound to the Rashba spin splitting in this material.

Since the measured trends we observe for T1 indicate dominant precessional spin relaxation, being inconsistent with all other spin relaxation mechanisms, our findings provide an unambiguous experimental detection of Rashba coupling in the 2D perovskite structure investigated here. The precessional spin relaxation mechanism has not been identified previously in the hybrid organic-inorganic perovskites,43,44 despite the fact that this mechanism dominates the room temperature spin dynamics in the majority of III-V semiconductors.35 An experimental verification that the physics of spin relaxation in traditional III-V’s semiconductors can be transferred to the perovskite’s is highly significant because III-V materials have been the primary focus for spintronic device development. The value we estimate from our theoretical simulations for the spin splitting (10 meV at an excess energy of 50 meV above the band gap) is 20-fold larger than GaAs48 but is in line with other inorganic systems possessing large spin-orbit coupling that are strong contenders for the realization of practical spintronic devices including GaSb quantum wells42 and InAs/GaSb superlattices.37,52,53 Our findings therefore highlight the promise of perovskite materials for applications in semiconductor spintronics, offering a solution-processable alternative for facile, low-cost device fabrication.

While the value of the Rashba coupling parameter we extract (λR = 0.08 eV Å) is a lower bound due to the influence of carrier thermalization as discussed above, this value is smaller than some recent reports of Rashba effects in the organic inorganic perovskites.61,68 The value of the spin splitting in the family of 2D perovskites will be dictated by the net distortions of the lead-iodide octahedra induced by competition between the long organic cations and the small cations.24 Varying the type of long organic cation while keeping the number of lead-iodide octahedra layers constant would therefore provide a useful extension to the present work by probing the extent to which the strength of Rashba coupling may be engineered in these systems. The first observation of the Rashba effect in an organic-inorganic perovskite (in the valence band of a single crystal sample of the 3D perovskite CH3NH3PbBr3) was carried out using ARPES,61 yielding a record Rashba coupling parameter of 7 eV Å. Recent reports suggest that this large Rashba effect was due to inversion asymmetry breaking at the surface.62–65 Such a surface Rashba effect is consistent with the observation of strong Rashba coupling in thin films of 3D CH3NH3PbI3 using the photogalvanic effect.66 In another study using ARPES on a similar single crystal sample, however, no k-dependent spin splitting was detected,81 indicating that the sample preparation conditions may influence the measured results.63 Strong Rashba coupling (λR = 1.6 eV Å) was recently found using electro-absorption techniques and density functional theory (DFT) calculations in the 2D perovskite (PEA)2PbI4 (PEPI, with a single layer of lead-iodide octahedra).68 In that measurement, the Rashba splitting was determined from a low-energy peak in the electro-absorption spectrum attributed to transitions between the Rashba spin split bands.68 Signatures of dynamic Rashba effects were also seen using transient polarization anisotropy in thin films of 3D CH3NH3PbI3.67 While the large size of the Rashba coupling observed in the perovskites in recent years is promising for spintronic applications, the wide spread of reported values highlights the need for further experimental studies of Rashba coupling using a range of experimental techniques. Our findings illustrate the utility of spin lifetime measurements for studying such effects, building upon the extensive existing literature probing precessional spin relaxation in inorganic semiconductors.

We have measured the spin relaxation time for optically injected carriers in thin films of the 2D perovskite (BA)2MAPb2I7 using time-resolved circular-dichroism techniques. The measured dependence of the spin lifetime on the laser excess energy and excited carrier density indicate precessional spin relaxation caused by the Rashba effect. Simulations of the measured carrier spin dynamics incorporating LO phonon and electron-electron scattering enabled the extraction of a Rashba spin splitting of 10 meV at an electron energy of 50 meV above the band gap. Our experiments, which confirm the presence of a Rashba spin splitting experimentally in a 2D perovskite material, open the door to studies of spin relaxation kinetics as a probe of Rashba effects in a wide range of 2D and 3D perovskite materials. For instance, the extension to larger thicknesses of the inorganic layer [e.g., (BA)2(MA)n−1PbnI3n+1 with n > 2] would enable an interesting comparison to the quantum well-width dependence in III-V semiconductors,82 and spin lifetime studies may be used to complement structural characterization of the temperature-dependent distortions of the metal halide octahedra within each crystal phase in the perovskites.83,83 Our findings shed-light on the spin-related properties of the organic-inorganic perovskites and point to the potential for spintronic devices based on these materials.

See supplementary material for details of sample preparation, linear absorption and x-ray diffraction, description of pump probe techniques, method of extraction of spin lifetimes from measured circular dichroism, and additional details of theoretical model including calculated scattering rates for electrons.

Work at Dalhousie University was supported by the Natural Sciences and Engineering Research Council of Canada. Work at Northwestern University was supported by Grant No. SC0012541 from the US Department of Energy, Office of Science. Work at Washington State University was supported by Grant No. W911NF-17-1-0511 from the US Army Research Office.

1.
See http://www.nrel.gov/ncpv/images/efficiency_chart.jpg for the National Renewable Energy Laboratory (NREL) chart of current record solar cell efficiencies.
2.
K.
Tanaka
,
T.
Takahashi
,
T.
Ban
,
T.
Kondo
,
K.
Uchida
, and
N.
Miura
, “
Comparative study on the excitons in lead-halide-based perovskite-type crystals CH3NH3PbBr3 CH3NH3PbI3
,”
Solid State Commun.
127
,
619
623
(
2003
).
3.
M.
Hirasawa
,
T.
Ishihara
,
T.
Goto
,
K.
Uchida
, and
N.
Miura
, “
Magnetoabsorption of the lowest exciton in perovskite-type compound (CH3NH3) PbI3
,”
Physica B
201
,
427
430
(
1994
).
4.
A.
Miyata
,
A.
Mitioglu
,
P.
Plochocka
,
O.
Portugall
,
J. T.-W.
Wang
,
S. D.
Stranks
,
H. J.
Snaith
, and
R. J.
Nicholas
, “
Direct measurement of the exciton binding energy and effective masses for charge carriers in organic-inorganic tri-halide perovskites
,”
Nat. Phys.
11
,
582
587
(
2015
).
5.
V.
D’Innocenzo
,
G.
Grancini
,
M. J. P.
Alcocer
,
A. R. S.
Kandada
,
S. D.
Stranks
,
M. M.
Lee
,
G.
Lanzani
,
H. J.
Snaith
, and
A.
Petrozza
, “
Excitons versus free charges in organo-lead tri-halide perovskites
,”
Nat. Commun.
5
,
3586
(
2014
).
6.
Y.
Yamada
,
T.
Nakamura
,
M.
Endo
,
A.
Wakamiya
, and
Y.
Kanemitsu
, “
Photoelectronic responses in solution-processed perovskite CH3NH3PbI3 solar cells studied by photoluminescence and photoabsorption spectroscopy
,”
IEEE J. Photovoltaics
5
,
401
405
(
2015
).
7.
D. A.
Valverde-Chávez
,
C. S.
Ponseca
, Jr.
,
C. C.
Stoumpos
,
A.
Yartsev
,
M. G.
Kanatzidis
,
V.
Sundström
, and
D. G.
Cooke
, “
Intrinsic femtosecond charge generation dynamics in single crystal CH3NH3PbI3
,”
Energy Environ. Sci.
8
,
3700
3707
(
2015
).
8.
H.-H.
Fang
,
R.
Raissa
,
M.
Abdu-Aguye
,
S.
Adjokatse
,
G. R.
Blake
,
J.
Even
, and
M. A.
Loi
, “
Photophysics of organic-inorganic hybrid lead iodide perovskite single crystals
,”
Adv. Funct. Mater.
25
,
2378
2385
(
2015
).
9.
F.
Thouin
,
S.
Neutzner
,
D.
Cortecchia
,
V. A.
Dragomir
,
C.
Soci
,
T.
Salim
,
Y. M.
Lam
,
R.
Leonelli
,
A.
Petrozza
,
A. R. S.
Kandada
, and
C.
Silva
, “
Stable biexcitons in two-dimensional metal-halide perovskites with strong dynamic lattice disorder
,”
Phys. Rev. Mater.
2
,
034001
(
2018
).
10.
S. A.
March
,
C.
Clegg
,
D. B.
Riley
,
D.
Webber
,
I. G.
Hill
, and
K. C.
Hall
, “
Simultaneous observation of free and defect-bound excitons in CH3NH3PbI3 using four-wave mixing spectroscopy
,”
Sci. Rep.
6
,
39139
(
2016
).
11.
S. A.
March
,
D. B.
Riley
,
C.
Clegg
,
D.
Webber
,
X.
Liu
,
M.
Dobrowolska
,
J. K.
Furdyna
,
I. G.
Hill
, and
K. C.
Hall
, “
Four-wave mixing in perovskite photovoltaic materials reveals long dephasing times and weaker many-body interactions than GaAs
,”
ACS Photonics
4
,
1515
1521
(
2017
).
12.
D.
Webber
,
C.
Clegg
,
A. W.
Mason
,
S. A.
March
,
I. G.
Hill
, and
K. C.
Hall
, “
Carrier diffusion in thin-film CH3NH3PbI3 perovskite measured using four-wave mixing
,”
Appl. Phys. Lett.
111
,
121905
(
2017
).
13.
J. S.
Manser
,
J. A.
Christians
, and
P. V.
Kamat
, “
Intriguing optoelectronic properties of metal halide perovskites
,”
Chem. Rev.
116
,
12956
13008
(
2016
).
14.
T. C.
Sum
and
N.
Mathews
, “
Advancements in perovskite solar cells: Photophysics behind the photovoltaics
,”
Energy Environ. Sci.
7
,
2518
2534
(
2014
).
15.
L. M.
Herz
, “
Charge-carrier dynamics in organic-inorganic metal halide perovksites
,”
Annu. Rev. Phys. Chem.
67
,
65
89
(
2016
).
16.
S. D.
Stranks
,
G. E.
Eperon
,
G.
Grancini
,
C.
Menelaou
,
M. J. P.
Alcocer
,
T.
Leijtens
,
L. M.
Herz
,
A.
Petrozza
, and
H. J.
Snaith
, “
Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber
,”
Science
342
,
341
344
(
2013
).
17.
G.
Xing
,
N.
Mathews
,
S.
Sun
,
S. S.
Lim
,
Y. M.
Lam
,
M.
Gratzel
,
S.
Mhaisalkar
, and
T. C.
Sum
, “
Long-range balanced electron- and hole-transport lengths in organic-inorganic CH3NH3PbI3
,”
Science
342
,
344
347
(
2013
).
18.
C. C.
Stoumpos
,
C. D.
Malliakas
, and
M. G.
Kanatzidis
, “
Semiconducting tin and lead iodide perovskites with organic cations: Phase transitions, high mobilities, and near-infrared photoluminescent properties
,”
Inorg. Chem.
52
,
9019
9038
(
2013
).
19.
M. H.
Kumar
,
S.
Dharani
,
W. L.
Leong
,
P. P.
Boix
,
R. R.
Prabhakar
,
T.
Baikaie
,
C.
Shi
,
H.
Ding
,
R.
Ramesh
,
M.
Asta
,
M.
Graetzel
,
S. G.
Mhaisalkar
, and
N.
Mathews
, “
Lead-free halide perovskite solar cells with high photocurrents realized though vacancy modulation
,”
Adv. Mater.
26
,
7122
7127
(
2014
).
20.
K. P.
Marshall
,
R. I.
Walton
, and
R. A.
Hatton
, “
Tin perovskite/fullerene planar layer photovoltaics: Improving the efficiency and stability of lead-free devices
,”
J. Mater. Chem. A
3
,
11631
11640
(
2015
).
21.
W.
Ke
,
C. C.
Stoumpos
,
M.
Zhu
,
L.
Mao
,
I.
Spanopoulos
,
J.
Liu
,
O. Y.
Kontsevoi
,
M.
Chen
,
D.
Sarma
,
Y.
Zhang
,
M. R.
Wasielewski
, and
M. G.
Kanatzidis
, “
Enhanced photovoltaic performance and stability with a new type of hollow 3D perovskite [en] FASnI3
,”
Sci. Adv.
3
,
e1701293
(
2017
).
22.
B.
Saparov
,
F.
Hong
,
J.-P.
Sun
,
H.-S.
Duan
,
W.
Meng
,
S.
Cameron
,
I. G.
Hill
,
Y.
Yan
, and
D. B.
Mitzi
, “
Thin-film preparation and characterization of Cs3Sb2I9: A lead-free layered perovskite semiconductor
,”
Chem. Mater.
27
,
5622
5632
(
2015
).
23.
D. H.
Cao
,
C. C.
Stoumpos
,
O. K.
Farha
,
J. T.
Hupp
, and
M. G.
Kanatzidis
, “
2D homologous perovskites as light-absorbing materials for solar cell applications
,”
J. Am. Chem. Soc.
137
,
7843
7850
(
2015
).
24.
C. C.
Stoumpos
,
D. H.
Cao
,
D. J.
Clark
,
J.
Young
,
J. M.
Rondinelli
,
J. I.
Jang
,
J. T.
Hupp
, and
M. G.
Kanatzidis
, “
Ruddlesden-Popper hybrid lead iodide perovskite 2D homologous semiconductors
,”
Chem. Mater.
28
,
2852
2867
(
2016
).
25.
I. C.
Smith
,
E. T.
Hoke
,
D.
Solis-Ibarra
,
M. D.
McGehee
, and
H. I.
Karunadasa
, “
A layered hybrid perovskite solar-cell absorber with enhanced moisture stability
,”
Angew. Chem.
126
,
11414
11417
(
2014
).
26.
C. R.
Kagan
,
D. B.
Mitzi
, and
C. D.
Dimitrakopoulos
, “
Organic-inorganic hybrid materials as semiconducting channels in thin-film field effect transistors
,”
Science
286
,
945
947
(
1999
).
27.
X. Y.
Chin
,
D.
Cortecchia
,
J.
Yin
,
A.
Bruno
, and
C.
Soci
, “
Lead iodide perovskite light-emitting field-effect transistor
,”
Nat. Commun.
6
,
7383
(
2015
).
28.
K.
Shibuya
,
M.
Koshimizu
,
Y.
Takeoka
, and
K.
Asai
, “
Scintillation properties of (C6H13NH3)2PbI4: Exciton luminescence of an organic/inorganic multiple quantum well structure compound induced by 2.0 MeV protons
,”
Nucl. Instrum. Methods Phys. Res., Sect. B
194
,
207
212
(
2002
).
29.
T.
Ishihara
,
J.
Takahashi
, and
T.
Goto
, “
Exciton state in two-dimensional perovskite semiconductor (C10H21NH3)2PbI4
,”
Solid State Commun.
69
,
933
936
(
1989
).
30.
Z.-K.
Tan
,
R. S.
Moghaddamn
,
M. L.
Lai
,
P.
Docampo
,
R.
Higler
,
F.
Deschler
,
M.
Price
,
A.
Sadhanala
,
L. M.
Pazos
,
D.
Credgington
,
F.
Hanusch
,
T.
Bein
,
H. J.
Snaith
, and
R. H.
Friend
, “
Bright light-emitting diodes based on organometal halide perovskite
,”
Nat. Nanotechnol.
9
,
687
692
(
2014
).
31.
F.
Yan
,
J.
Xing
,
G.
Xing
,
L. N.
Quan
,
S. T.
Tan
,
J.
Zhao
,
R.
Su
,
L.
Zhang
,
S.
Chen
,
Y.
Zhao
,
A.
Huan
,
E. H.
Sargent
,
Q.
Xiong
, and
H. V.
Demir
, “
Highly efficient visible colloidal lead-halide perovskite nanocrystal light-emitting diodes
,”
Nano Lett.
18
,
3157
3164
(
2018
).
32.
H.
Zhu
,
Y.
Fu
,
F.
Meng
,
X.
Wu
,
Z.
Gong
,
Q.
Ding
,
M. V.
Gustafsson
,
M. T.
Trinh
,
S.
Jin
, and
X. Y.
Zhu
, “
Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors
,”
Nat. Mater.
14
,
636
642
(
2015
).
33.
L.
Dou
,
Y.
Yang
,
J.
You
,
Z.
Hong
,
W.-H.
Chang
,
G.
Li
, and
Y.
Yang
, “
Solution-processed hybrid perovskite photodetectors with high detectivity
,”
Nat. Commun.
5
,
5404
(
2014
).
34.
M.
Kepenekian
and
J.
Even
, “
Rashba and Dresselhaus couplings in halide perovskites: Accomplishments and opportunities for spintronics and spin-orbitronics
,”
J. Phys. Chem. Lett.
8
,
3362
3370
(
2017
).
35.
Semiconductor Spintronics and Quantum Computation
, edited by
D. D.
Awschalom
,
D.
Loss
, and
N.
Samarth
(
Springer-Verlag
,
Berlin Heidelberg, New York
,
2002
).
36.
S.
Datta
and
B.
Das
, “
Electronic analog of the electro-optic modulator
,”
Appl. Phys. Lett.
56
,
665
667
(
1990
).
37.
K. C.
Hall
and
M. E.
Flatté
, “
Performance of a spin-based insulated gate field effect transistor
,”
Appl. Phys. Lett.
88
,
162503
(
2006
).
38.
K.
Onodera
,
T.
Masumoto
, and
M.
Kimura
, “
980 nm compact optical isolators using Cd1−x−yMnxHgyTe single crystals for high power pumping laser diodes
,”
Electron. Lett.
30
,
1954
1955
(
1994
).
39.
J.
Rudolph
,
D.
Hagele
,
H. M.
Gibbs
,
G.
Khitrova
, and
M.
Oestreich
, “
Laser threshold reduction in a spintronic device
,”
Appl. Phys. Lett.
82
,
4516
4518
(
2003
).
40.
K. C.
Hall
,
J. P.
Zahn
,
A.
Gamouras
,
S.
March
,
J. L.
Robb
,
X.
Liu
, and
J. K.
Furdyna
, “
Ultrafast optical control of coercivity in GaMnAs
,”
Appl. Phys. Lett.
93
,
032504
(
2008
).
41.
Y.
Nishikawa
,
A.
Tackeuchi
,
S.
Nakamura
,
S.
Muto
, and
N.
Yokoyama
, “
All-optical picosecond switching of a quantum well etalon using spin-polarization relaxation
,”
Appl. Phys. Lett.
66
,
839
841
(
1995
).
42.
K. C.
Hall
,
S. W.
Leonard
,
H. M.
van Driel
,
A. R.
Kost
,
E.
Selvig
, and
D. H.
Chow
, “
Subpicosecond spin relaxation in GaAsSb multiple quantum wells
,”
Appl. Phys. Lett.
75
,
4156
4158
(
1999
).
43.
D.
Giovanni
,
H.
Ma
,
J.
Chua
,
M.
Gratzel
,
R.
Ramesh
,
S.
Mhaisalkar
,
N.
Mathews
, and
T. C.
Sum
, “
Highly spin-polarized carrier dynamics and ultralarge photoinduced magnetization in CH3NH3PbI3 perovskite thin films
,”
Nano Lett.
15
,
1553
1558
(
2015
).
44.
P.
Odenthal
,
W.
Talmadge
,
N.
Gundlach
,
R.
Wang
,
C.
Zhang
,
D.
Sun
,
Z.-G.
Yu
,
Z. V.
Vardeny
, and
Y. S.
Li
, “
Spin-polarized exciton quantum beating in hybrid organic-inorganic perovskites
,”
Nat. Phys.
13
,
894
899
(
2017
).
45.
D.
Giovanni
,
W. K.
Chong
,
H. A.
Dewi
,
K.
Thirumal
,
I.
Neogi
,
R.
Ramesh
,
S.
Mhaisalkar
,
N.
Mathews
, and
T. C.
Sum
, “
Tunable room-temperature spin-selective optical Stark effect in solution-processed layered halide perovskites
,”
Sci. Adv.
2
,
e1600477
(
2016
).
46.
J.
Even
,
L.
Pedesseau
,
M.-A.
Dupertuis
,
J.-M.
Jancu
, and
C.
Katan
, “
Electronic model for self-assembled hybrid organic/perovskite semiconductors: Reverse band edge electronic states ordering and spin-orbit coupling
,”
Phys. Rev. B
86
,
205301
(
2012
).
47.
J.
Even
,
L.
Pedesseau
,
J.-M.
Jancu
, and
C.
Katan
, “
Importance of spin-orbit coupling in hybrid organic/inorganic perovskites for photovoltaic applications
,”
J. Phys. Chem. Lett.
4
,
2999
3005
(
2013
).
48.
W. H.
Lau
,
J. T.
Olesberg
, and
M. E.
Flatte
, “
Electron-spin decoherence in bulk and quantum-well zinc-blende semiconductors
,”
Phys. Rev. B
64
,
161301(R)
(
2001
).
49.
G.
Dresselhaus
, “
Spin-orbit coupling effects in zinc blende structures
,”
Phys. Rev.
100
,
580
586
(
1955
).
50.
Y. A.
Bychkov
and
E. I.
Rashba
, “
Properties of a 2D electron gas with lifted spectral degeneracy
,”
JETP Lett.
39
,
78
81
(
1984
).
51.
P.
Pfeffer
and
W.
Zawadzki
, “
Spin splitting of conduction subbands in III-V heterostructures due to inversion asymmetry
,”
Phys. Rev. B
59
,
R5312
R5315
(
1999
).
52.
K. C.
Hall
,
K.
Gundogdu
,
E.
Altunkaya
,
W. H.
Lau
,
M. E.
Flatte
,
T. F.
Boggess
,
J. J.
Zinck
,
W. B.
Barvosa-Carter
, and
S. L.
Skeith
, “
Spin relaxation in (110) and (001) InAs/GaSb superlattices
,”
Phys. Rev. B
68
,
115311
(
2003
).
53.
K. C.
Hall
,
W. H.
Lau
,
K.
Gundogdu
,
M. E.
Flatte
, and
T. F.
Boggess
, “
Nonmagnetic semiconductor spin transistor
,”
Appl. Phys. Lett.
83
,
2937
2939
(
2003
).
54.
F.
Brivio
,
K. T.
Butler
,
A.
Walsh
, and
M.
van Schilfgaarde
, “
Relativistic quasiparticle self-consistent electronic structure of hybrid halide perovskite photovoltaic absorbers
,”
Phys. Rev. B
89
,
155204
(
2014
).
55.
A.
Amat
,
E.
Mosconi
,
E.
Ronca
,
C.
Quarti
,
P.
Umari
,
M. K.
Nazeeruddin
,
M.
Gratzel
, and
F.
De Angelis
, “
Cation-induced band-gap tuning in organohalide perovskites: Interplay of spin-orbit coupling and octahedra tilting
,”
Nano Lett.
14
,
3608
3616
(
2014
).
56.
A.
Stroppa
,
D.
Di Sante
,
P.
Barone
,
M.
Bokdam
,
G.
Kresse
,
C.
Franchini
,
M.-H.
Whangbo
, and
S.
Picozzi
, “
Tunable ferroelectric polarization and its interplay with spin-orbit coupling in tin iodide perovskites
,”
Nat. Commun.
5
,
5900
(
2014
).
57.
M.
Kim
,
J.
Im
,
A. J.
Freeman
,
J.
Ihm
, and
H.
Jin
, “
Switchable s = 1/2 and j = 1/2 Rashba bands in ferroelectric halide perovskites
,”
Proc. Natl. Acad. Sci. U. S. A.
111
,
6900
6904
(
2014
).
58.
M.
Kepenekian
,
R.
Robles
,
C.
Katan
,
D.
Sapori
,
L.
Pedesseau
, and
J.
Even
, “
Rashba and Dresselhaus effects in hybrid organic-inorganic perovskites: From basics to devices
,”
ACS Nano
9
,
11557
11567
(
2015
).
59.
L.
Leppert
,
S. E.
Reyes-Lillo
, and
J. B.
Neaton
, “
Electric field- and strain-induced Rashba effect in hybrid halide perovskites
,”
J. Phys. Chem. Lett.
7
,
3683
3689
(
2016
).
60.
Z.-G.
Yu
, “
Rashba effect and carrier mobility in hybrid organic-inorganic perovskites
,”
J. Phys. Chem. Lett.
7
,
3078
3083
(
2016
).
61.
D.
Niesner
,
M.
Wilhelm
,
I.
Levchuk
,
A.
Osvet
,
S.
Shrestha
,
M.
Batentschuk
,
C.
Brabec
, and
T.
Fauster
, “
Giant Rashba splitting in CH3NH3PbBr3 organic-inorganic perovskite
,”
Phys. Rev. Lett.
117
,
126401
(
2016
).
62.
E.
Mosconi
,
T.
Etienne
, and
F.
De Angelis
, “
Rashba band splitting in organohalide lead perovskites: Bulk and surface effects
,”
J. Phys. Chem. Lett.
8
,
2247
(
2017
).
63.
X.
Che
,
B.
Traore
,
C.
Katan
,
M.
Kepenekian
, and
J.
Even
, “
Does Rashba splitting in CH3NH3PbBr3 arise from 2×2 surface construction?
,”
Phys. Chem. Chem. Phys.
20
,
9638
9643
(
2018
).
64.
V.
Sarritzu
,
N.
Sestu
,
D.
Marongiu
,
X.
Chang
,
Q.
Wang
,
S.
Masi
,
S.
Colella
,
A.
Rizzo
,
A.
Gocalinska
,
E.
Pelucchi
,
M. L.
Mercuri
,
F.
Quochi
,
M.
Saba
,
A.
Mura
, and
G.
Bongiovanni
, “
Direct or indirect bandgap in hybrid lead halide perovskites?
,”
Adv. Opt. Mater.
6
,
1701254
(
2018
).
65.
K.
Frohna
,
T.
Deshpande
,
J.
Harter
,
W.
Peng
,
B. A.
Barker
,
J. B.
Neaton
,
S. G.
Louie
,
O. M.
Bakr
,
D.
Hsieh
, and
M.
Bernardi
, “
Inversion symmetry and bulk Rashba effect in methylammonium lead iodide perovskite single crystals
,”
Nat. Commun.
9
,
1829
(
2018
).
66.
D.
Niesner
,
M.
Hauck
,
S.
Shrestha
,
I.
Levchuk
,
G. J.
Matt
,
A.
Osvet
,
M.
Batentschuk
,
C.
Brabec
,
H. B.
Weber
, and
T.
Fauster
, “
Structural fluctuations cause spin-split states in tetragonal (CH3NH3) PbI3 as evidenced by the circular photogalvanic effect
,”
Proc. Natl. Acad. Sci. U. S. A.
115
,
9509
(
2018
).
67.
J. P. H.
Rivett
,
L. Z.
Tan
,
M. B.
Price
,
S. A.
Bourelle
,
N. J. L. K.
Davis
,
J.
Xiao
,
Y.
Zou
,
R.
Middleton
,
B.
Sun
,
A. M.
Rappe
,
D.
Credgington
, and
F.
Deschler
, “
Long-lived polarization memory in the electronic states of lead-halide perovskites from local structural dynamics
,”
Nat. Commun.
9
,
3531
(
2018
).
68.
Y.
Zhai
,
S.
Baniya
,
C.
Zhang
,
J.
Li
,
P.
Haney
,
C.-X.
Sheng
,
E.
Ehrenfreund
, and
Z. V.
Vardeny
, “
Giant Rashba splitting in 2D organic-inorganic halide perovskites measured by transient spectroscopies
,”
Sci. Adv.
3
,
e1700704
(
2017
).
69.
M. I.
D’yakonov
and
V. I.
Perel
, “
Spin relaxation of conduction electrons in noncentrosymmetric semiconductors
,”
Sov. Phys. Solid State
13
,
3023
3026
(
1972
).
70.
M. I.
D’yakonov
and
V. Yu.
Kachorovskii
, “
Spin relaxation of two-dimensional electrons in noncentrosymmetric semiconductors
,”
Sov. Phys. Semicond.
20
,
110
112
(
1986
).
71.

We note that the symmetry breaking tied to the crystal fields inherent to the crystal structure itself, for example due to polar bonding, is conventionally referred to as bulk inversion asymmetry. The x-ray diffraction measurements indicated that the symmetry breaking direction tied to the crystal structure in the 2D perovskite studied here is along the stacking direction, consistent with the C2v space group.

72.
R.
Terauchi
,
Y.
Ohno
,
T.
Adachi
,
A.
Sato
,
F.
Matsukura
,
A.
Tackeuchi
, and
H.
Ohno
, “
Carrier mobility dependence of electron spin relaxation in GaAs quantum wells
,”
Jpn. J. Appl. Phys., Part 1
38
(
1
),
2549
2551
(
1999
).
73.
R. L.
Milot
,
R. J.
Sutton
,
G. E.
Eperon
,
A. A.
Haghighirad
,
J. M.
Hardigree
,
L.
Miranda
,
H. J.
Snaith
,
M. B.
Johnston
, and
L. M.
Herz
, “
Charge-carrier dynamics in 2D hybrid metal-halide perovskites
,”
Nano Lett.
16
,
7001
7007
(
2016
).
74.

The coherent artifact is caused by four-wave mixing on the probed optical transitions and is only present for temporally overlapped pump and probe pulses.

75.
D.
Wang
,
B.
Wen
,
Y.-N.
Zhu
,
C.-J.
Tong
,
Z.-K.
Tang
, and
L.-M.
Liu
, “
First-principles study of novel two-dimensional (C4H9NH3)2PbX4 perovskites for solar cell absorbers
,”
J. Phys. Chem. Lett.
8
,
876
883
(
2017
).
76.
R. J.
Elliott
, “
Theory of the effect of spin-orbit coupling on magnetic resonance in some semiconductors
,”
Phys. Rev.
96
,
266
279
(
1954
).
77.
Y.
Yafet
, “
g-factors and spin-lattice relaxation of conduction electrons
,”
Solid State Phys.
14
,
1
98
(
1963
).
78.
G. L.
Bir
,
A. G.
Aronov
, and
G. E.
Pikus
, “
Spin relaxation of electrons due to scattering by holes
,”
Sov. Phys. JETP
42
,
705
712
(
1976
).
79.
L.
Ni
,
U.
Huynh
,
A.
Cheminal
,
T. H.
Thomas
,
R.
Shivanna
,
T. F.
Hinrichsen
,
S.
Ahmad
,
A.
Sadhanala
, and
A.
Rao
, “
Real-time observation of exciton-phonon coupling dynamics in self-assembled hybrid perovskite quantum wells
,”
ACS Nano
11
,
10834
10843
(
2017
).
80.
T.
Sohier
,
M.
Calandra
, and
F.
Mauri
, “
Two-dimensional Fröhlich interaction in transition-metal dichalcogenide monolayers: Theoretical modeling and first-principles calculations
,”
Phys. Rev. B
94
,
085415
(
2016
).
81.
T.
Komesu
,
X.
Huang
,
T. R.
Paudel
,
Y. B.
Losovyj
,
X.
Zhang
,
E. F.
Schwier
,
Y.
Kojima
,
M.
Zheng
,
H.
Iwasawa
,
K.
Shimada
,
M. I.
Saidaminov
,
D.
Shi
,
L.
Abdelhady
,
O. M.
Bakr
,
S.
Dong
,
E. Y.
Tsymbal
, and
P. A.
Dowben
, “
Surface electronic structure of hybrid organo lead bromide perovskite single crystals
,”
J. Phys. Chem. Lett.
120
,
21710
21715
(
2016
).
82.
A.
Tackeuchi
,
O.
Wada
, and
Y.
Nishikawa
, “
Electron spin relaxation in InGaAs/InP multiple-quantum wells
,”
Appl. Phys. Lett.
70
,
1131
1133
(
1997
).
83.
D. G.
Billing
and
A.
Lemmerer
, “
Synthesis, characterization and phase transitions in the inorganic-organic layered perovskite-type hybrids [(CnH2n+1NH3)2PbI4], n = 4, 5, and 6
,”
Acta Crystallogr., Sect. B: Struct. Sci.
63
,
735
747
(
2007
).

Supplementary Material