This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Topologically-enhanced exciton transport

Joshua J. P. Thompson Department of Materials Science and Metallurgy, University of Cambridge, 27 Charles Babbage Road, Cambridge CB3 0FS, United Kingdom    Wojciech J. Jankowski Theory of Condensed Matter Group, Cavendish Laboratory, University of Cambridge, J. J. Thomson Avenue, Cambridge CB3 0HE, United Kingdom    Robert-Jan Slager Department of Physics and Astronomy, The University of Manchester, Oxford Road, Manchester M13 9PL, United Kingdom Theory of Condensed Matter Group, Cavendish Laboratory, University of Cambridge, J. J. Thomson Avenue, Cambridge CB3 0HE, United Kingdom    Bartomeu Monserrat Department of Materials Science and Metallurgy, University of Cambridge, 27 Charles Babbage Road, Cambridge CB3 0FS, United Kingdom
(August 6, 2025)
Abstract

Excitons dominate the optoelectronic response of many materials. Depending on the time scale and host material, excitons can exhibit free diffusion, phonon-limited diffusion, or polaronic diffusion, and exciton transport often limits the efficiency of optoelectronic devices such as solar cells or photodetectors. We demonstrate that topological excitons exhibit enhanced diffusion in all transport regimes. Using quantum geometry, we find that topological excitons are generically larger and more dispersive than their trivial counterparts, promoting their diffusion. We apply this general theory to organic polyacene semiconductors and show that exciton transport increases up to fourfold when topological excitons are present. We also propose that non-uniform electric fields can be used to directly probe the quantum metric of excitons, providing a rare experimental window into a basic geometric feature of quantum states. Our results provide a new strategy to enhance exciton transport in semiconductors and reveal that mathematical ideas of topology and quantum geometry can be important ingredients in the design of next-generation optoelectronic technologies.

I Introduction

Excitons, Coulomb-bound electron-hole pairs, dominate the optoelectronic response of a multitude of semiconductors Wang et al. (2018); Perea-Causin et al. (2022); Posmyk et al. (2024). Prominent examples include organic Mikhnenko et al. (2015); Valencia et al. (2023) and low-dimensional Chernikov et al. (2014); Yu et al. (2015); Dresselhaus et al. (2007) semiconductors, each a vast and versatile family of compounds which host excitons with large binding energies that can reach hundreds of milielectronvolts Mikhnenko et al. (2015); Giannini et al. (2022). The formation, dynamics, lifetime, and transport of excitons dictate the efficiency of a host of technological applications, from solar cells Wilson et al. (2011); Congreve et al. (2013) and light-emitting diodes Xu et al. (2021); Chowdhury et al. (2025), to biosensors Shanmugaraj and John (2019); Geldert et al. (2017). From a material perspective, the chemical and structural diversity available in the design of organic and low-dimensional semiconductors allows fine-tuning of the electronic and excitonic properties for customized device applications Huang et al. (2018).

Despite their promise, one of the key limitations of organic semiconductors is the low mobility of excitons Giannini et al. (2022); Sneyd et al. (2021); Muth et al. (2024). For example, low exciton mobility has been shown to inhibit the efficiency of organic semiconductor based solar cells Ghorab et al. (2022) since excitons decay before being extracted. As another example, in some organic systems the fission of optically active singlet excitons into pairs of optically inactive triplet excitons could help boost efficiency beyond the Shockley-Queisser limit Congreve et al. (2013), but the diffusion of these triplets is even slower than that of singlets Muth et al. (2024), and again exciton transport is a limiting factor. Other schemes, such as organic co-crystals Gunder et al. (2021); Chen et al. (2023) and organic-inorganic interfaces Bettis Homan et al. (2017); Bowman et al. (2022); Thompson et al. (2023a), again suffer from exciton mobility limitations.

Refer to caption
Figure 1: Topologically-enhanced diffusive transport of topological excitons with inversion symmetry-protected topological 2\mathbb{Z}_{2}-invariant PexcP_{\text{exc}} in the presence of phonons (wiggly lines). Due to exciton-phonon interactions, the propagating excitons in topological excitonic band can be scattered, dephased, and the diffusive transport of the excitons can be further altered with non-uniform electric fields introducing a controllable forcing. We show that non-trivial excitonic quantum geometry can be manifested in all these transport features.

In this work, we propose topology as a new avenue to enhance exciton transport. The topology of electrons is well-established Qi and Zhang (2011); Hasan and Kane (2010); Armitage et al. (2018), leading to remarkable transport properties such as quantum Hall phenomena Chang et al. (2023). A natural question to ask is whether topological ideas can be extended to excitons, which are starting to be explored in two-dimensional van der Waals layered materials Wu et al. (2017); Kwan et al. (2021), organic semiconductors Jankowski et al. (2024), and idealised models Davenport et al. (2024); Zhu et al. (2024). In this context, a recent remarkable result concerns the topologically-induced non-trivial Riemannian geometry of exciton wavefunctions. Specifically, it has been demonstrated that exciton quantum geometry, phrased in terms of the quantum metric Provost and Vallee (1980), provides a lower bound on the centre-of-mass spread ξ\xi of excitons Jankowski et al. (2024):

ξ2a2Pexc24,\xi^{2}\geq\frac{a^{2}P_{\text{exc}}^{2}}{4}, (1)

where aa is the lattice parameter of the crystal and PexcP_{\text{exc}} is an excitonic topological invariant protected by crystalline inversion symmetry. The relationship between topology, quantum geometry, and exciton properties is general and can be applied to different topological invariants in different dimensions and in different material platforms Bouhon et al. (2023). For example, in a one-dimensional setting, exemplified by organic polyacenes Cirera et al. (2020); Romanin et al. (2022); Jankowski et al. (2024), the excitonic topology can be characterised through a topological invariant Pexc2P_{\text{exc}}\in\mathbb{Z}_{2}, associated with the first Stiefel-Whitney characteristic class w12w_{1}\in\mathbb{Z}_{2} that reflects the unorientability of an excitonic band Jankowski et al. (2024); Ahn et al. (2019); Bouhon et al. (2020).

Qualitatively, the bound in Eq. (1) implies that topological excitons are more delocalised, and can be larger, than their trivial counterparts. In this work, we exploit this key insight to demonstrate that topological excitons exhibit enhanced transport compared to their trivial counterparts. We demonstrate enhanced exciton transport in all regimes, ranging from free exciton diffusion at femtosecond timescales to phonon-limited and polaronic diffusion at longer picosecond timescales. We also illustrate these general results in a family of organic polyacene crystals, where we find that topological excitons exhibit a four-fold increase in their transport compared to their trivial counterparts. Overall, our work establishes topology as a new avenue for improving optoelectronic technologies.

II Topological excitons: model and materials

To explore the role of topology on exciton transport, we focus on a one-dimensional system that has recently been predicted to host topological excitons Jankowski et al. (2024). In this setting, single-particle electron properties are described by the Su-Schrieffer-Heeger (SSH) model Su et al. (1979, 1980):

H=t1jcB,jcA,jt2jcB+1,jcA,j+h.c.,\begin{split}H=-t_{1}\sum_{j}c^{\dagger}_{B,j}c_{A,j}-t_{2}\sum_{j}c^{\dagger}_{{B+1},j}c_{A,j}+\text{h.c.},\end{split} (2)

with cB,j/cA,jc^{\dagger}_{B,j}/c_{A,j} being the creation/annihilation operators for the electrons at sublattices A,BA,B, in unit cell jj, and alternating hopping parameters t1t_{1} and t2t_{2}. The topological phase realising topological edge states corresponds to t2>t1t_{2}>t_{1}, and the trivial phase corresponds to t2<t1t_{2}<t_{1}Su et al. (1979, 1980).

From these single-particle electron and hole states, we then describe the exciton properties using the Wannier equation Thompson et al. (2023b, a); Jankowski et al. (2024), which directly incorporates the electron-hole Coulomb interaction. The solution of the Wannier equation yields exciton bands EνQE_{\nu Q} associated with exciton states |ψνQexc\ket{\psi^{\text{exc}}_{\nu Q}}, where ν\nu is the band index and QQ is the exciton centre-of-mass momentum. We also introduce |uνQexc\ket{u^{\text{exc}}_{\nu Q}} as the cell-periodic part of the excitonic Bloch state |ψνQexc=eiQR|uνQexc\ket{\psi^{\text{exc}}_{\nu Q}}=e^{iQR}\ket{u^{\text{exc}}_{\nu Q}}, where R=(re+rh)/2R=(r_{\mathrm{e}}+r_{\mathrm{h}})/2 is the centre-of-mass position of the exciton, with electron position rer_{\mathrm{e}} and hole position rhr_{\mathrm{h}}. The topology of excitons in one-dimensional centrosymmetric semiconductors can be captured by a 2\mathbb{Z}_{2} invariant PexcP_{\text{exc}}, which can be directly obtained from the excitonic states Jankowski et al. (2024).

A material realisation of this model is provided by polyacene chains composed of nn-ring acene molecules, where n=3,5,7n=3,5,7, linked by a carbon-carbon bond on the central carbon atoms. Illustrative examples, polyanthracene (n=3n=3) and polypentacence (n=5n=5), are shown in Fig. 1. These polyacenes exhibit a topologically trivial exciton phase with Pexc=0P_{\text{exc}}=0 for n=3n=3, and a topological phase with Pexc=1P_{\text{exc}}=1 for n=5,7n=5,7Jankowski et al. (2024).

We emphasise that the model and materials described above are for illustrative purposes only, and the key findings of this work are generally applicable to the transport of topological excitons in any material and dimension.

III Free Exciton Propagation

Upon photoexcitation, excitons diffuse freely at femtosecond timescales Rosati et al. (2020); Ashoka et al. (2022); Zhang et al. (2022). The exciton diffusion constant is given by (see SM):

Dν=122EνQQ2+1μνΔQμνgxxμν(Q),D_{\nu}=\frac{1}{2\hbar}\left\langle\frac{\partial^{2}E_{\nu Q}}{\partial Q^{2}}\right\rangle+\frac{1}{\hbar}\sum_{\mu\neq\nu}\langle\Delta^{\mu\nu}_{Q}g^{\mu\nu}_{xx}(Q)\rangle, (3)

where EνQE_{\nu Q} is the exciton energy dispersion for band ν\nu, ΔQμν=Eμ(Q)Eν(Q)\Delta^{\mu\nu}_{Q}=E_{\mu}(Q)-E_{\nu}(Q) is the energy difference between the pair of exciton bands μ\mu and ν\nu, and gxxμν(Q)=QuμQexc|uνQexcuνQexc|QuμQexcg^{\mu\nu}_{xx}(Q)=\braket{\partial_{Q}u^{\text{exc}}_{\mu Q}|u^{\text{exc}}_{\nu Q}}\braket{u^{\text{exc}}_{\nu Q}|\partial_{Q}u^{\text{exc}}_{\mu Q}} is the excitonic multiband quantum metric. The exciton diffusivity in Eq. (3) has two contributions: the first term arises from the energy dispersion, and the second term arises from the quantum geometric properties of the associated exciton states.

We next show that the geometric term in the exciton diffusivity of Eq. (3) leads to enhanced transport for topological excitons. Starting with the flat band limit, the contribution from the exciton energy dispersion vanishes, as 2EνQ/Q2=0\partial^{2}E_{\nu Q}/\partial Q^{2}=0. Therefore, the exciton diffusion comes entirely from the geometric contribution. The geometric contribution scales according to ΔQμνgxxμν(Q)1/ΔQμν\Delta^{\mu\nu}_{Q}g^{\mu\nu}_{xx}(Q)\propto 1/\Delta^{\mu\nu}_{Q} (see SM), and in the flat band limit, focusing on the lowest exciton band, we can approximate the geometric contribution to the diffusivity as Dν=1Δμν=1gxxμν(Q){D_{\nu=1}\approx\frac{\Delta}{\hbar}\sum_{\mu\neq\nu=1}\langle g^{\mu\nu}_{xx}(Q)\rangle}, where Δ\Delta is the smallest QQ-independent gap from the band. We can then define the Brillouin zone average quantum metric gxxν(Q)\langle g^{\nu}_{xx}(Q)\rangle associated with exciton band ν\nu by tracing over the interband contributions according to gxxν(Q)=μνgxxμν(Q)\langle g^{\nu}_{xx}(Q)\rangle=\sum_{\mu\neq\nu}\langle g^{\mu\nu}_{xx}(Q)\rangle, and we obtain that the diffusivity in the flat band limit is:

Dν=1Δgxxν=1(Q).D_{\nu=1}\approx\frac{\Delta}{\hbar}\langle g^{\nu=1}_{xx}(Q)\rangle. (4)

Using the bound ξ2a2Pexc24\xi^{2}\geq\frac{a^{2}P_{\text{exc}}^{2}}{4} from Eq. (1), and noting that the exciton centre-of-mass spread is related to the metric according to ξ2=gxxν=1(Q)\xi^{2}=\langle g^{\nu=1}_{xx}(Q)\rangle, we identify a lower bound on the geometric contribution to the exciton diffusivity:

Dν=1Δa2Pexc24.D_{\nu=1}\geq\frac{\Delta}{\hbar}\frac{a^{2}P_{\text{exc}}^{2}}{4}. (5)

Therefore, diffusive exciton transport in the lowest exciton band is directly impacted by the underlying exciton topology: the geometric contribution to the exciton diffusivity exhibits a lower bound for topological excitons (Pexc=1)P_{\mathrm{exc}}=1) but no bound for trivial excitons (Pexc=0P_{\mathrm{exc}}=0). This is a direct consequence of the lower bound on the exciton centre-of-mass spread, as determined by quantum geometry, which makes topological excitons larger and therefore facilitates diffusion.

Moving to the general dispersive case, both topological and trivial excitons will have equivalent contributions from the band dispersion to the diffusivity. Therefore, we can generally claim that topological excitons in the diffusive regime exhibit enhanced transport compared to their trivial counterparts.

To numerically illustrate the above results, we consider polypentacene as an example of a material hosting topological excitons with t2>t1t_{2}>t_{1}. We construct an initial exciton wavepacket, formed around the photoexcitation spot, and we calculate the subsequent exciton diffusion that leads to the spatial spread depicted in Fig. 2(a). To test the importance of the topology-bound geometric contribution, we also consider the scenario in which the values of the hopping parameters are swapped, so that t2<t1t_{2}<t_{1} and we are in the trivial regime. The corresponding exciton diffusion is depicted in Fig. 2(b). The two scenarios have the same band dispersion, leading to the same first term in Eq. (3). However, the topological exciton diffuses more rapidly, a consequence of the geometric term in the diffusivity, which is bounded by below for topological excitons. These results explicitly demonstrate that exciton diffusion is enhanced in polypentacene driven by the underlying exciton topology.

More generally, Fig. 2(c) presents the diffusion constant as a function of t2t_{2} and t1t_{1} allowing us to demonstrate the wide applicability of our results. For any pair {t1,t2}\{t_{1},t_{2}\}, if t2>t1t_{2}>t_{1} (topological excitons) then the diffusion constant is significantly larger than for the equivalent pair with t1>t2t_{1}>t_{2} (trivial). When t1t_{1} and t2t_{2} are significantly different, the resulting diffusivities can differ by several orders of magnitude. The band contribution to the diffusion for topological and trivial excitons is equivalent [Fig. 2(d)], peaking at t1t2t_{1}\sim t_{2}, where the electron and exciton band structures become most dispersive. In contrast the geometric contribution is distinctly larger for the topological excitons compared to trivial ones [Fig. 2(e)]. Non-zero topological transport in the flat band limit can be seen by comparing the t1=0t_{1}=0 or t2=0t_{2}=0 limit in Fig. 2(d,e) for the topological and trivial excitons, respectively.

Refer to caption
Figure 2: Time-dependent transport of topologically non-trivial (a) and trivial (b) excitons. The diffusivity of topologically non-trivial excitons is bounded from below by the excitonic 2\mathbb{Z}_{2} invariant. Parameters for n=5n=5 polypentacene are used, with the DFT predicted combination of intracell (t1t_{1}) and intercell (t2t_{2}) hoppings employed in (a) while the order is flipped in (b), to directly ascertain the impact of topology. (c) Exciton diffusion constant as a function of t1t_{1} and t2t_{2} with t2>t1t_{2}>t_{1} (t2<t1t_{2}<t_{1}) representing topological and trivial excitons respectively Jankowski et al. (2024). Breakdown of the contribution to the exciton diffusion, shown in (c), of exciton dispersion (d) and exciton geometry (e).
Refer to caption
Figure 3: (a) Schematic of non-uniform electric field on polypentacene crystal. The quantum metric of the topological exciton leads to a larger force due to the electric field compared to the trivial case. (b) Electric field induced tuning of the exciton group velocity v~Q\tilde{v}_{Q} of the lowest exciton band for different values of t2t_{2} and fixed t1=0.33t_{1}=0.33 eV. Our extracted value of t2t_{2} for polypentacene from DFT is 0.520.52 eV. The solid coloured (dashed black) lines show the exciton dispersions with (without) an applied electric field. The velocity plots with increasing t1t_{1} are offset by 0.5 eV for clarity.

IV Exciton transport in non-uniform electric fields

We next explore driven exciton transport under non-uniform electric fields, which we demonstrate can be used to directly probe the exciton quantum geometry. The exciton group velocity vνQ\langle v_{\nu Q}\rangle associated with band ν\nu is given in one dimension by (see Methods) :

vνQ=vνQ0μνe22Q(gxxμν(Q)ΔQμν)(rR(R))2,\langle v_{\nu Q}\rangle=\langle v^{0}_{\nu Q}\rangle-\sum_{\mu\neq\nu}\frac{e^{2}}{\hbar^{2}}\partial_{Q}\Bigg{(}\frac{g^{\mu\nu}_{xx}(Q)}{\Delta^{\mu\nu}_{Q}}\Bigg{)}\Big{(}\langle r\rangle\cdot\nabla_{R}\mathcal{E}(R)\Big{)}^{2}, (6)

where vνQ0\langle v^{0}_{\nu Q}\rangle is the free exciton group velocity and R(R)\nabla_{R}\mathcal{E}(R) is the applied electric field gradient which couples to the electron-hole distance r\langle r\rangle. According to Eq. (6), the total exciton group velocity has a contribution from the free exciton group velocity vνQ0\langle v^{0}_{\nu Q}\rangle, and a contribution from the quantum metric derivatives. In one dimension, the latter can be described by the Christoffel symbols Γxxxμν(Q)=12Qgxxμν(Q)\Gamma^{\mu\nu}_{xxx}(Q)=\frac{1}{2}\partial_{Q}g^{\mu\nu}_{xx}(Q). Overall, an exciton moving in a non-uniform electric field experiences a force, leading to either acceleration or deceleration of the exciton, and a modulation of the exciton group velocity.

The geometric contribution to the exciton group velocity in Eq. (6) depends on the energy difference ΔQμν\Delta^{\mu\nu}_{Q} between bands μ\mu and ν\nu. This dependence can be suppressed by increasing dielectric screening, for example through strongly polar substrates, such that ΔQμνΔ\Delta^{\mu\nu}_{Q}\approx\Delta can be made approximately uniform over the exciton Brillouin zone. In this regime, the non-linear exciton transport in non-uniform electric fields is directly given by the quantum geometric Christoffel symbols.

In Fig. 3(a), we schematically show the impact of an applied non-uniform electric field on exciton transport, where topological excitons experience an enhanced transport. Quantitatively, we perform a numerical simulation of the exciton group velocity for polypentacene, and additional simulations where we vary t2t_{2} freely, and for simplicity we set the electric field gradient to be constant R(R)=0.1\nabla_{R}\mathcal{E}(R)=0.1 V/nm2. Figure 3(b) shows the excitonic group velocity vνQ\langle v_{\nu Q}\rangle modulated by a non-uniform electric field (coloured, shaded) at different values of t2t_{2} for a fixed value t1t_{1}= 0.3 eV. The group velocity vQv_{Q} in the absence of an external field is shown with the dashed lines. In the trivial regime, vνQvνQ0\langle v_{\nu Q}\rangle\approx\langle v^{0}_{\nu Q}\rangle due to the vanishing quantum metric gxxμν0g_{xx}^{\mu\nu}\approx 0 and vanishing variations thereof, Γxxxμν0\Gamma^{\mu\nu}_{xxx}\approx 0. The topological regime (t2>t1t_{2}>t_{1}) shows a more complex behaviour. At small finite QQ, the exciton diffusion is slowed down by the electric field with vνQvνQ0\langle v_{\nu Q}\rangle\ll\langle v^{0}_{\nu Q}\rangle and even shows an opposite sign. At larger QQ, the force induced by the non-uniform electric field on the topological excitons becomes larger, leading to a huge enhancement of the excitonic group velocity. This effect is most significant for t2t_{2} reasonably close to t1t_{1} within the range t2<0.5t_{2}<0.5 eV. For larger t2t_{2}, the quantum metric contribution shrinks, owing to the smaller excitons Jankowski et al. (2024) such that the group velocity with and without electric field begin to converge again, see the red curve Fig. 3(b).

Qualitatively, the distinct response of topological and trivial excitons under a non-uniform electric field can again be related to their different centre-of-mass localisations and relative sizes. Trivial excitons have a smaller size, and therefore are less subject to electric field gradients. The quantum metric in the momenta conjugate to the relative electron-hole position rr, and the centre-of-mass coordinates RR, precisely reflect the corresponding spreads and localisations of excitons (see Methods).

V Phonon-limited exciton diffusion

Following free exciton diffusion at femtosecond timescales, excitons experience phonon-limited diffusion at picosecond timescales Thompson et al. (2022); Cohen et al. (2024). In this regime, the exciton diffusion is given by:

Dph=Q,νvνQ2ΓνQeEνQ/kBT𝒵,D_{\text{ph}}=\sum_{Q,\nu}\frac{\langle v_{\nu Q}^{2}\rangle}{\Gamma_{\nu Q}}\frac{e^{-E_{\nu Q}/k_{\mathrm{B}}T}}{\mathcal{Z}}, (7)

where vνQv_{\nu Q} is the exciton group velocity, ΓνQ\Gamma_{\nu Q} is the exciton-phonon scattering rate, EνQE_{\nu Q} is the exciton band energy, kBk_{\mathrm{B}} is the Boltzmann constant, TT is temperature, and 𝒵\mathcal{Z} is the partition function. The role that topology and quantum geometry play on phonon-limited exciton diffusion depends on the interplay between the exciton group velocity and exciton-phonon scattering rates featuring in Eq. (7).

Starting with the exciton group velocity, topological excitons exhibit enhanced group velocities (see Methods):

vμQ2=vμQ2+12νμ|ΔQμν|2gxxμν(Q),\langle v^{2}_{\mu Q}\rangle=\langle{v}_{\mu Q}\rangle^{2}+\frac{1}{\hbar^{2}}\sum_{\nu\neq\mu}|\Delta^{\mu\nu}_{Q}|^{2}g^{\mu\nu}_{xx}(Q), (8)

where the second term represents the geometric contribution that enhances the group velocity of topological excitons.

In terms of exciton-phonon scattering rates, the key microscopic quantities are the exciton-phonon scattering matrix elements 𝒟Qqβμν\mathcal{D}^{\mu\nu}_{Qq\beta} which describe the scattering from an initial exciton (ν,Q)(\nu,Q) into a final exciton (μ,Q+q)(\mu,Q+q) mediated by a phonon (β,q)(\beta,q) of momentum qq and energy ωβq\hbar\omega_{\beta q}. In turn, the exciton-phonon matrix elements can be written in terms of individual electron-phonon scattering matrix elements gkqβmng^{mn}_{kq\beta} modulated by the exciton envelope function (see Methods). The electron-phonon scattering matrix elements describe the scattering from an initial electron (hole) (n,k)(n,k) into a final electron (hole) (m,k+q)(m,k+q) mediated by a phonon (β,q)(\beta,q). Topological electrons were previously found to significantly contribute to the electron-phonon coupling underpinned by gkqβmng^{mn}_{kq\beta} through electronic quantum geometric terms Yu et al. (2024). As a consequence, topological electrons enhance exciton-phonon coupling matrix elements 𝒟Qqβμν\mathcal{D}^{\mu\nu}_{Qq\beta}, and we confirm this numerically as shown in Fig. 4(a-b).

The preceding discussion implies that topological electrons will enhance the resulting exciton-phonon scattering matrix elements, but not all topological electrons lead to topological excitons. Topological excitons can arise from obstructed electrons and holes Jankowski et al. (2024), and in this scenario the topology-enhanced electron-phonon scattering matrix elements will result in topology-enhaned exciton-phonon matrix elements. These in turn will lead to enhanced exciton-phonon scattering rates ΓνQ\Gamma_{\nu Q}. However, unobstructed electrons and holes can also give rise to topological excitons due to the electron-hole contribution Jankowski et al. (2024); Davenport et al. (2024). In this second scenario, there is no enhancement of the electron-phonon scattering matrix elements, resulting in topological excitons that exhibit no enhancement in the exciton-phonon scattering rates ΓνQ\Gamma_{\nu Q}.

Overall, we end up with two scenarios. In the first scenario, the diffusion of topological excitons is enhanced when the underlying electrons and holes are trivial, driven by the topologically-driven enhancement of the exciton group velocity vνQv_{\nu Q}. In the second scenario, corresponding to topological excitons with underlying topological electrons and holes, both the exciton group velocity vνQv_{\nu Q} and the exciton-phonon scattering rates ΓνQ\Gamma_{\nu Q} are enhanced. The diffusivity of Eq.(7) depends on the ratio vνQ/ΓνQv_{\nu Q}/\Gamma_{\nu Q}, and therefore the diffusion of topological excitons in this scenario may be enhanced or suppressed. In the numerical example below, the enhancement of the group velocity dominates and the topological excitons exhibit enhanced transport.

To illustrate these results numerically, we consider the topological excitons in polyacenes. Polyacenes exhibit topological excitons with underlying topological electrons and holes. This is the only regime we can explore as there are no known material candidates hosting topological excitons with underlying trivial electrons and holes. In Fig. 4(a-b) we show the exciton-phonon scattering matrix elements from an initial state QQ to a final state QQ^{\prime} for polypentacene (t2>t1t_{2}>t_{1}) and compare it to the trivial counterpart where the values of the hopping parameters are swapped (t2<t1t_{2}<t_{1}). We use dimensionless units, as we are interested in the impact of the topology rather than the absolute values of the matrix elements. We observe different couplings for different momenta, depending on the topology associated with the Zak phases of the electronic and hole states comprising the excitons, with the peak intensities being dictated by the quantum geometry of individual electrons and holes, as well as their momentum-dependent interaction. We find that the Q/QQ/Q^{\prime} dependence on the exciton-phonon coupling is the same in both the trivial and topological case, however the magnitude is significantly enhanced in the topological case as expected from the discussion above.

Refer to caption
Figure 4: Topology-dependent exciton-phonon coupling in polypentacene. (a-b) Exciton-phonon matrix elements resolved in initial QQ and final QQ^{\prime} excitonic centre-of-mass momentum. (c) Phonon-induced exciton dephasing as function of initial momentum QQ at 50 K (orange) and 300 K (blue). The trivial and topological exciton dephasings are shown by the solid and dashed lines respectively. (d) Phonon-stimulated exciton diffusion for trivial (blue) and topological (pink) excitons.

One way to probe the impact of phonon scattering is via the exciton dephasing ΓQ=νΓνQ\Gamma_{Q}=\sum_{\nu}\Gamma_{\nu Q}. When Q=0Q=0, the dephasing corresponds to the non-radiative lifetime of the lowest exciton state. In Fig. 4(c), we present the calculated exciton dephasing as a function of momentum at 5050 K (orange) and 300300 K (blue) for polypentacene (solid lines) and its trivial counterpart (dashed lines). The dephasing depends on the population of phonons, which increases as a function of temperature. As such the dephasing at 300300 K is significantly larger than that at 5050 K. Irrespective of temperature, the dephasing is larger in the topological case, which can be understood by the larger magnitude of the exciton-phonon matrix elements of the topological regime compared to the trivial one. The excitonic dispersions themselves are almost identical, so any density of states effects in the allowed scattering channels Thompson et al. (2022) are approximately equivalent for both trivial and topological exciton dephasing. As a result, the same qualitative features are observed in the dephasing curves for both topological and trivial exictons at high and low temperatures. An initial increase in the dephasing can be observed at small QQ, characterised by the emission of acoustic phonons scattering back to the Q=0Q=0 state or at larger temperatures, absorption of phonons. This gives rise to a distinct bump feature Thompson et al. (2022) between Q=0.1Q=0.1 nm-1 and Q=0.5Q=0.5 nm-1. At around Q=0.8Q=0.8 nm-1, the exciton energy difference compared to Q=0Q=0 nm-1 corresponds to the optical phonon energy. As a result intraband optical phonon relaxation becomes possible leading to a sharp increase in the dephasing. At larger momentum QQ such relaxation remains possible, however, the exciton density of states at higher-momentum states is lower, leading to an overall decrease in the dephasing. At very large Q>3Q>3 nm-1, the exciton band flattens (cosine-like) leading to an increase in the excitonic density of states and a corresponding increase in scattering channels. As a result, peak is seen in the 300 K dephasing at Q=4.7Q=4.7 nm-1, but an equivalent peak is not present in the 50 K results as the thermal occupation of optical phonons is very small in the latter case.

We also calculate the phonon-limited exciton diffusion coefficients using Eq. (7) and report the results in Fig. 4(d). In this example, the competition between the geometric contribution to the excitonic group velocity and to the enhanced exciton-phonon coupling leads to an overall enhancement of the diffusion in the topological regime. Taking solely the band contribution to the exciton velocity, the increased exciton-phonon dephasing associated with topological excitons leads to topological excitons diffusing about four times more slowly than trivial excitons at all temperatures, see Fig. S1. However, taking the exciton band geometry into account, leads to an increase in the exciton group velocity at low QQ in both the trivial and topological regime. While a fairly modest increase in the case of trivial excitons, the vastly enhanced exciton metric in the topological regime leads to a large increase in the exciton group velocity of the low QQ yet highly populated Q=0Q=0 states. As a result the exciton diffusion is much larger in the topological regime, even despite the enhanced exciton-phonon coupling which increases the scattering term. The temperature dependence in Fig. 4(d) reflects this interpretation, with low temperatures corresponding to an increase in the relative population of low momentum excitons which have a large group velocity enhancement. The reduced exciton-phonon coupling at low temperatures adds to this behaviour and we see a monotonic decrease in the exciton diffusion in both trivial and topological excitons.

VI Polaronic effects

When the interaction between excitons and phonons becomes sufficiently large, excitons can become localised by the lattice Markvart and Greef (2004); Hansen et al. (2022), becoming heavier and undergoing slower transport. These exciton-polarons have been studied extensively Baranowski et al. (2024); Knorr et al. (2024); Dai et al. (2024a), and are particularly relevant in organic systems where their formation hinders the already limited energy transfer across organic optoelectronic devices Coehoorn et al. (2017). Hence, understanding the transport of excitons in this strong coupling regime is crucial.

We calculate the exciton band dispersion for polypentacene as renormalised by exciton-polarons at 300300 K by treating the exciton-phonon interaction self-consistently (see Methods). The renormalised exciton-polaron dispersion for polypentacene is shown in Fig. 5(a), with the trivial counterpart shown in Fig. 5(b). In both cases, polaron formation results in an energy shift and in a decrease in the group velocity, but notably the topological exciton-polaron exhibits a larger energy shift and a larger reduction in velocity with a correspondingly increased mass, which we attribute to the topology-driven increase in the exciton-phonon interactions. Figure  5(d) shows the ratio of the free excitonic to the polaronic group velocities. We find the usual low-momentum decrease of the exciton-polaron velocity (vExc/vExc-Pol>1v_{\text{Exc}}/v_{\text{Exc-Pol}}>1 ) in both the trivial (green) and topological (red) cases, corresponding to a polaron velocity around 70%70\% of that of the free exciton.

Experimentally, the formation of exciton polarons will lead to a red shift of the excitonic resonance energy Hurtado Parra et al. (2022) in the spectrum of absorbed/emitted light on a picosecond timescale. Importantly, the band topology of the exciton-polarons is the same as that of the bare excitons, and we note that at large QQ the topological exciton-polaron bands are close in energy but do not cross, analogous to the bare exciton case. Our results show that the band contribution to the polaron velocity is reduced in the topological case, however the same metric contribution to the exciton transport holds, given that the polaron bands do not cross and possess the same underlying metric.

Refer to caption
Figure 5: Transport of topological exciton-polarons. Exciton-polaron and exciton band dispersion for (a) topological and (b) trivial excitons. A clear polaron shift is observed in both cases.(c) Schematic of reduced transport of exciton-polarons compared to bare excitons. (d) Ratio of free exciton to exciton-polaron group velocities in polypentacene at 300300 K for trivial (green) and topological (red) regimes. The blue region indicates mass enhancement and slower exciton-polarons while the pink region indicates mass reduction and faster excitons.

VII Discussion

Overall, our results show that the topologically-bounded localisation properties of excitons dramatically affect their transport properties. Compared to their trivial counterparts, topological excitons sustain faster free transport, have lower effective masses, and their transport signatures are more robust to polaronic effects. The enhanced transport of topological excitons is expected to be experimentally trackable in the polyacenes Cirera et al. (2020), where the underlying electronic topology has already been observed. Experimentally, the dynamics of exciton transport can be visualised with time-resolved photoluminescence Sung et al. (2019); Rosati et al. (2020) or with transient absorption Schnedermann et al. (2019).

We show that topological excitons also experience stronger electron-phonon coupling-driven exciton-phonon coupling, compared to their trivial counterparts. This observation respects the expected enhancement of electron-phonon coupling of the constituent electrons and holes that host non-trivial quantum geometry Yu et al. (2024). As discussed earlier, the stronger exciton-phonon coupling experienced by topological excitons results in higher dephasing rates, but we find that these are not sufficient for the topological excitons to violate the original quantum geometric bounds of the free exciton propagation, as compared to the trivial excitons. Similarly, the transport of topological excitons in the polaronic regime is also robust against coupling to the lattice of organic crystals. These findings, accounting for the presence of physical effects present in all semiconducting materials, show that our diagnosis of quantum geometric manifestations on excitons should persist under experimental conditions.

Finally, we stress that the exciton transport properties and the associated exciton quantum geometry and topology can be controlled using an appropriate dielectric environment Jankowski et al. (2024), chemical modifications, and temperature, which modifies the population of the exciton (and phonon) states. Therefore, our findings provide a general quantum mechanical formalism and mathematical insights to theoretically understand the experimentally controllable geometric manifestations due to excitonic topologies, as reflected in the discussed excitonic transport in semiconductor materials.

VIII Conclusions

We have demonstrated that the transport of topological excitons is significantly enhanced compared to that of trivial excitons. This discovery arises from the lower bound that the centre-of-mass excitonic quantum geometry sets on the exciton localisation, making topological excitons larger and therefore more mobile. We have shown that enhanced topological exciton transport holds in sub-picosecond free transport regime and in the picosecond phonon-limited and polaronic transport regimes. Additionally, we have illustrated these discoveries in a family of polyacene organic semiconductors. Our results are general, and we expect that exciton topology can be exploited to enhance the transport properties of a wide variety of semiconductors for applications in optoelectronic devices.


Acknowledgements.
The authors thank Richard Friend and Akshay Rao for helpful discussions. J.J.P.T. and B.M. acknowledge support from a EPSRC Programme Grant [EP/W017091/1]. W.J.J. acknowledges funding from the Rod Smallwood Studentship at Trinity College, Cambridge. R.-J.S. acknowledges funding from a New Investigator Award, EPSRC grant EP/W00187X/1, a EPSRC ERC underwrite grant EP/X025829/1, a Royal Society exchange grant IES/R1/221060, and Trinity College, Cambridge. B.M. also acknowledges support from a UKRI Future Leaders Fellowship [MR/V023926/1] and from the Gianna Angelopoulos Programme for Science, Technology, and Innovation. Calculations were performed using the Sulis Tier-2 HPC platform hosted by the Scientific Computing Research Technology Platform at the University of Warwick. Sulis is funded by EPSRC Grant [EP/T022108/1] and the HPC Midlands+ consortium.

IX Methods

IX.1 Exciton quantum geometry

We consider an exciton state associated with exciton band ν\nu and centre-of-mass momentum Q:

|ψνQexc=𝐤ψν𝐐(𝐤)ei𝐤𝐫|u𝐤+𝐐/2e|u𝐤+𝐐/2h,\ket{\psi^{\text{exc}}_{\nu\textbf{Q}}}=\sum_{\bf k}\psi_{\nu{\bf Q}}({\bf k})e^{i{\bf k}\cdot{\bf r}}\ket{u^{\text{e}}_{\mathbf{k+Q}/2}}\ket{u^{\text{h}}_{\mathbf{-k+Q}/2}}, (9)

where ψν𝐐(𝐤)\psi_{\nu{\bf Q}}({\bf k}) is the envelope function capturing the electron-hole correlation, 𝐫=𝐫e𝐫h{\bf r}={\bf r}_{\text{e}}-{\bf r}_{\text{h}} is the relative electron-hole distance with the associated relative momentum 𝐤{\bf k}, and |u𝐤+𝐐/2e\ket{u^{\text{e}}_{\mathbf{k+Q}/2}} and |u𝐤+𝐐/2h\ket{u^{\text{h}}_{\mathbf{-k+Q}/2}} are the single-particle electron and hole states. Exploiting translational symmetry, we can also write the exciton state as:

|ψνQexc=ei𝐐𝐑|uνQexc,\ket{\psi^{\text{exc}}_{\nu\textbf{Q}}}=e^{\text{i}{\bf Q}\cdot{\bf R}}\ket{u^{\text{exc}}_{\nu\textbf{Q}}}, (10)

where 𝐑=𝐫e+𝐫h2\mathbf{R}=\frac{\mathbf{r}_{\text{e}}+\mathbf{r}_{\text{h}}}{2} is the center-of-mass coordinate, and the exciton state satisfies Bloch’s theorem with the cell-periodic part given by:

|uνQexc=ei𝐐𝐑𝐤ei𝐤𝐫ψν𝐐(𝐤)|u𝐤+𝐐/2e|u𝐤+𝐐/2hei𝐐𝐑𝐤ei𝐤𝐫|uνQ,kexc.\ket{u^{\text{exc}}_{\nu\textbf{Q}}}=e^{-\text{i}{\bf Q}\cdot{\bf R}}\sum_{\bf k}e^{\text{i}{\bf k}\cdot{\bf r}}\psi_{\nu{\bf Q}}({\bf k})\ket{u^{\text{e}}_{\mathbf{k+Q}/2}}\ket{u^{\text{h}}_{\mathbf{-k+Q}/2}}\equiv e^{-\text{i}{\bf Q}\cdot{\bf R}}\sum_{\bf k}e^{\text{i}{\bf k}\cdot{\bf r}}\ket{u^{\text{exc}}_{\nu\textbf{Q},\textbf{k}}}. (11)

The quantum geometry associated with exciton states was originally introduced in Ref. Jankowski et al. (2024). The quantum geometric tensor in the centre-of-mass coordinates (𝐑)iiQi(\mathbf{R})_{i}\sim\text{i}\partial_{Q_{i}} is given by:

𝒬ijexc,ν(Q)=QiuνQexc|(1P^ν𝐐)|QjuνQexc,\mathcal{Q}^{\text{exc},\nu}_{ij}(\textbf{Q})=\bra{\partial_{Q_{i}}u^{\text{exc}}_{\nu\textbf{Q}}}(1-\hat{P}_{\nu{\bf Q}})\ket{\partial_{Q_{j}}u^{\text{exc}}_{\nu\textbf{Q}}}, (12)

where P^ν𝐐=|uνQexcuνQexc|\hat{P}_{\nu{\bf Q}}=\ket{u^{\text{exc}}_{\nu\textbf{Q}}}\bra{u^{\text{exc}}_{\nu\textbf{Q}}} is a projector onto the exciton band of interest. Its real part, the quantum metric, is given by:

gijexc,ν(Q)=QiuνQexc|(1P^ν𝐐)|QjuνQexc+QjuνQexc|(1P^ν𝐐)|QiuνQexc2,g^{\text{exc},\nu}_{ij}(\textbf{Q})=\frac{\bra{\partial_{Q_{i}}u^{\text{exc}}_{\nu\textbf{Q}}}(1-\hat{P}_{\nu{\bf Q}})\ket{\partial_{Q_{j}}u^{\text{exc}}_{\nu\textbf{Q}}}+\bra{\partial_{Q_{j}}u^{\text{exc}}_{\nu\textbf{Q}}}(1-\hat{P}_{\nu{\bf Q}})\ket{\partial_{Q_{i}}u^{\text{exc}}_{\nu\textbf{Q}}}}{2}, (13)

and it relates to the centre-of-mass spread of excitons (𝐑𝐑)2\langle(\mathbf{R}-\langle\mathbf{R}\rangle)^{2}\rangle. Importantly, the relation between the exciton spread and the quantum metric can be exploited to reconstruct the exciton quantum metric in time-dependent transport experiments that involve freely propagating and driven excitons, as we show in the main text.

IX.2 Time-dependent free exciton diffusion

In the free diffusive regime, following Fick’s second law, the temporal and spatial evolution of the exciton density can be expressed as:

ρν(x,t)=N02π(2Dνt+σIni2)exp[(xxIni)22(2Dνt+σIni2)],\displaystyle\rho^{\nu}(x,t)=\dfrac{N_{0}}{\sqrt{2\pi(2D_{\nu}t+\sigma_{\text{Ini}}^{2})}}\exp\left[\dfrac{-(x-x_{\text{Ini}})^{2}}{2(2D_{\nu}t+\sigma_{\text{Ini}}^{2})}\right], (14)

where N0N_{0} is the initial number of generated excitons in excitonic band ν\nu, and for well-localised excitons we have σIni2ξ2\sigma_{\text{Ini}}^{2}\approx\xi^{2}, where xInix_{\text{Ini}} and σIni\sigma_{\text{Ini}} are the initial excitation centre and broadening, respectively.

In the following, we show that the exciton diffusivity DνD_{\nu} in band ν\nu is fully captured by the centre-of-mass quantum metric of the excitons gxx(Q)g_{xx}(Q). By mapping the quantum dynamics of free excitons to Fokker-Planck Gaussian propagation in one spatial dimension, we obtain that σ2(t)=σIni2+2Dνt{\sigma^{2}(t)=\sigma_{\text{Ini}}^{2}+2D_{\nu}t}, with Dν=2mνgxxν(Q)D_{\nu}=\frac{\hbar}{2m^{*}_{\nu}}\rightarrow\langle g^{\nu}_{xx}(Q)\rangle. The full derivation is detailed in the Supplemental Material (SM), but briefly, we map the density time-evolution equation to the Focker-Planck equation, in order to connect the diffusivity to the effective excitonic mass mνm^{*}_{\nu}. Furthermore, we utilise the Hellmann-Feynman theorem to derive the relation between the effective excitonic mass (mνm^{*}_{\nu}) and the quantum-geometry in the centre-of-mass momentum space. As a result, we find that the diffusivity of excitons in band ν\nu is given by:

Dν=122EνQQ2+1μνΔQμνgxxμν(Q),D_{\nu}=\frac{1}{2\hbar}\left\langle\frac{\partial^{2}E_{\nu Q}}{\partial Q^{2}}\right\rangle+\frac{1}{\hbar}\sum_{\mu\neq\nu}\langle\Delta^{\mu\nu}_{Q}g^{\mu\nu}_{xx}(Q)\rangle, (15)

where EνQE_{\nu Q} is a dispersion of band ν\nu, and the averages are taken with respect to the Brillouin zone spanned in the QQ momentum space parameter (see also SM).

IX.3 Driven exciton transport under non-uniform electric fields

We consider exciton transport driven by an external non-uniform electric field gradient, complementary to the field gradients realisable internally within the system Yang et al. (2016). Semiclassically, interacting electrons and holes satisfy the equation of motion Chaudhary et al. (2021):

𝐤˙e/h=𝐫e/hU(𝐫e𝐫h)e(𝐫e/h),\dot{{\bf k}}_{\text{e/h}}=-\nabla_{{\bf r}_{\text{e/h}}}U({\bf r}_{\text{e}}-{\bf r}_{\text{h}})\mp e\mathcal{E}({\bf r}_{\text{e/h}}), (16)

where U(𝐫e𝐫h)U({\bf r}_{\text{e}}-{\bf r}_{\text{h}}) is the electron-hole interaction potential. This implies that the centre-of-mass exciton momentum Q=𝐤e+𝐤h\textbf{Q}={\bf k}_{\text{e}}+{\bf k}_{\text{h}} satisfies an equation of motion with a position-dependent external force 𝐅(𝐑){\bf F}({\bf R}):

Q˙=e[(𝐫h)(𝐫e)]=e[(R+𝐫/2)(R𝐫/2)]=e𝐫R(R)=𝐅(𝐑).\dot{\textbf{Q}}=e[\mathcal{E}({\bf r}_{h})-\mathcal{E}({\bf r}_{e})]=e[\mathcal{E}(\textbf{R}+{\bf r}/2)-\mathcal{E}(\textbf{R}-{\bf r}/2)]=e\langle{\bf r}\rangle\cdot\nabla_{\textbf{R}}\mathcal{E}(\textbf{R})={\bf F}({\bf R}). (17)

In this expression, we use R=(𝐫e+𝐫h)/2\textbf{R}=({\bf r}_{\text{e}}+{\bf r}_{\text{h}})/2, 𝐫=𝐫e𝐫h{{\bf r}={\bf r}_{\text{e}}-{\bf r}_{\text{h}}} and that to first order, (R±𝐫/2)=(R)±𝐫/2R(R)+O(𝐫2)\mathcal{E}(\textbf{R}\pm{\bf r}/2)=\mathcal{E}(\textbf{R})\pm{\bf r}/2\cdot\nabla_{\textbf{R}}\mathcal{E}(\textbf{R})+O({\bf r}^{2}).

Physically, the quantum geometric coupling to Q˙\dot{\textbf{Q}} can be related to the renormalised exciton energies. Consider a perturbation coupling to the centre of mass of the exciton ΔH=𝐑𝐅(𝐑)\Delta H=-{\bf R}\cdot{\bf F}({\bf R}), where 𝐑{\bf R} is a position operator projected onto an excitonic band. For the off-diagonal elements, we have ψμ𝐐exc|𝐑|ψν𝐐exc=iuμ𝐐exc|𝐐uμ𝐐exc\bra{\psi^{\text{exc}}_{\mu{\bf Q}}}{\bf R}\ket{\psi^{\text{exc}}_{\nu{\bf Q}}}=\text{i}\braket{u^{\text{exc}}_{\mu{\bf Q}}|\nabla_{\bf Q}u^{\text{exc}}_{\mu{\bf Q}}}, whereas the diagonal elements vanish by parity. At second order in perturbation theory, and assuming that the exciton bands are non-degenerate, we obtain the following energy corrections:

E~ν𝐐=Eν𝐐μν|ψν𝐐exc|ΔH|ψμ𝐐exc|2Eμ𝐐Eν𝐐=Eν𝐐μν𝐅T(𝐑)ψν𝐐exc|𝐑|ψμ𝐐excψμ𝐐exc|𝐑|ψν𝐐exc𝐅(𝐑)Eμ𝐐Eν𝐐,\tilde{E}_{\nu{\bf Q}}=E_{\nu{\bf Q}}-\sum_{\mu\neq\nu}\frac{|\bra{\psi^{\text{exc}}_{\nu{\bf Q}}}\Delta H\ket{\psi^{\text{exc}}_{\mu{\bf Q}}}|^{2}}{E_{\mu{\bf Q}}-E_{\nu{\bf Q}}}=E_{\nu{\bf Q}}-\sum_{\mu\neq\nu}\frac{{\bf F}^{\text{T}}({\bf R})\cdot\bra{\psi^{\text{exc}}_{\nu{\bf Q}}}{\bf R}\ket{\psi^{\text{exc}}_{\mu{\bf Q}}}\bra{\psi^{\text{exc}}_{\mu{\bf Q}}}{\bf R}\ket{\psi^{\text{exc}}_{\nu{\bf Q}}}\cdot{\bf F}({\bf R})}{E_{\mu{\bf Q}}-E_{\nu{\bf Q}}}, (18)

which in terms of the excitonic quantum metric, we can rewrite as

E~νQ=EνQμνgxxμν(Q)EμQEνQF(R)F(R),\tilde{E}_{\nu Q}=E_{\nu Q}-\sum_{\mu\neq\nu}\frac{g^{\mu\nu}_{xx}(Q)}{E_{\mu Q}-E_{\nu Q}}F(R)F(R), (19)

for a one-dimensional system. Here, E~νQ\tilde{E}_{\nu Q} is the excitonic energy renormalised by the coupling to external force fields. Denoting ΔQμν=EμQEνQ\Delta^{\mu\nu}_{Q}=E_{\mu Q}-E_{\nu Q}, and substituting F(R)=Q˙=erR(R)F(R)=\hbar\dot{Q}=e\langle r\rangle\cdot\nabla_{R}\mathcal{E}(R), we arrive at:

vνQ=1QE~νQ=vνQ0μνe22Q(gxxμν(Q)ΔQμν)(rR(R))2,\langle v_{\nu Q}\rangle=\frac{1}{\hbar}\partial_{Q}\tilde{E}_{\nu Q}=\langle v^{0}_{\nu Q}\rangle-\sum_{\mu\neq\nu}\frac{e^{2}}{\hbar^{2}}\partial_{Q}\Bigg{(}\frac{g^{\mu\nu}_{xx}(Q)}{\Delta^{\mu\nu}_{Q}}\Bigg{)}\Big{(}\langle r\rangle\cdot\nabla_{R}\mathcal{E}(R)\Big{)}^{2}, (20)

where vνQ0=1QEνQ\langle v^{0}_{\nu Q}\rangle=\frac{1}{\hbar}\partial_{Q}E_{\nu Q} is the free exciton velocity.

The above result implies that varying the electric field gradient in transport experiments allows the reconstruction of the derivatives of the exciton quantum metric. As mentioned in the main text, in the flat-band limit ΔQμνΔ\Delta^{\mu\nu}_{Q}\approx\Delta, the Christoffel symbols Γxxxμν=12Qgxxμν(Q)\Gamma^{\mu\nu}_{xxx}=\frac{1}{2}\partial_{Q}g^{\mu\nu}_{xx}(Q) can be directly accessed with this strategy. It should be noted that the size of the exciton, given by the average of the relative electron-hole coordinate r\langle r\rangle, must be known to assess the magnitude of the force F(R)F(R) due to the electric field gradient R(R)\nabla_{R}\mathcal{E}(R). Correspondingly, we compute the average size of the exciton that is relevant for the semiclassical equation of motion directly from the envelope function: r=0drr×|ψνQ(r)|2\langle r\rangle=\int^{\infty}_{0}\mathop{}\!\mathrm{d}r~r\times|\psi_{\nu Q}(r)|^{2}, where ψνQ(r)\psi_{\nu Q}(r) is a Fourier transform of ψνQ(k)\psi_{\nu Q}(k)Jankowski et al. (2024).

From the perspective of quantum geometry, we note that the derivatives of the quantum metric defining the Christoffel symbols can be in principle arbitrarily high due to the envelope contributions to the excitonic quantum metric, resulting in a nearly step-like character for gxxexc(Q)g_{xx}^{\text{exc}}(Q) in the presence of a singular non-Abelian Berry connection. Such singular behaviours of non-Abelian excitonic Berry connection are only to be expected in topological excitonic phases, as in the trivial phases with vanishing topological invariants the Berry connection can be chosen to be globally smooth.

IX.4 Exciton group velocity

The group velocity term featuring in the phonon-limited exciton diffusion and in the exciton-polaron diffusion has a quantum geometric contribution. To derive it, we use a resolution of the identity in terms of excitonic states, 1=ν|uνQuνQ|1=\sum_{\nu}\ket{u_{\nu Q}}\bra{u_{\nu Q}} and find that:

vμQ2\displaystyle\langle v^{2}_{\mu Q}\rangle =12uμQ|(QHQ)2|uμQ\displaystyle=\frac{1}{\hbar^{2}}\bra{u_{\mu Q}}(\partial_{Q}H_{Q})^{2}\ket{u_{\mu Q}}
=12νuμQ|QHQ|uνQuνQ|QHQ|uμQ\displaystyle=\frac{1}{\hbar^{2}}\sum_{\nu}\bra{u_{\mu Q}}\partial_{Q}H_{Q}\ket{u_{\nu Q}}\bra{u_{\nu Q}}\partial_{Q}H_{Q}\ket{u_{\mu Q}}
=12(QEμQ)2+12νμuμQ|QH(Q)|uνQuνQ|QH(Q)|uμQ\displaystyle=\frac{1}{\hbar^{2}}\Big{(}\partial_{Q}E_{\mu Q}\Big{)}^{2}+\frac{1}{\hbar^{2}}\sum_{\nu\neq\mu}\bra{u_{\mu Q}}\partial_{Q}H(Q)\ket{u_{\nu Q}}\bra{u_{\nu Q}}\partial_{Q}H(Q)\ket{u_{\mu Q}}
=vμQ2+12νμ|ΔQμν|2gxxμν(Q),\displaystyle=\langle{v}_{\mu Q}\rangle^{2}+\frac{1}{\hbar^{2}}\sum_{\nu\neq\mu}|\Delta^{\mu\nu}_{Q}|^{2}g^{\mu\nu}_{xx}(Q), (21)

with ΔQμν=EμQEνQ\Delta^{\mu\nu}_{Q}=E_{\mu Q}-E_{\nu Q}, which allows the multiband exciton quantum metric elements gxxμν(Q)g^{\mu\nu}_{xx}(Q) to modify the phonon-mediated diffusion via interband velocity matrix elements. Intuitively, the latter determine the variance of the velocity operator. On substituting the exciton quantum metric-dependent vμQ2\langle v^{2}_{\mu Q}\rangle for DphD_{\text{ph}}, we observe that the geometric contribution enhances the phonon-mediated diffusion of the topological excitons.

IX.5 Exciton-phonon coupling

In this section, we consider the connection between exciton-phonon coupling (ExPC) matrix elements Antonius and Louie (2022) and quantum geometry. The electron-phonon coupling (EPC) Hamiltonian can be written as:

Hel-ph=k,m,n,q,βgkqβmn𝒂^mk+q𝒂^nk(𝒃^β,q+𝒃^β,q),\displaystyle H_{\text{el-ph}}=\sum_{k,m,n,q,\beta}g^{mn}_{kq\beta}\hat{\bm{a}}^{\dagger}_{mk+q}\hat{\bm{a}}_{nk}\left(\hat{\bm{b}}_{\beta,q}+\hat{\bm{b}}^{\dagger}_{\beta,-q}\right), (22)

where 𝒂^nk()\hat{\bm{a}}^{(\dagger)}_{nk} is the annihilation (creation) operator for an electron in band nn and momentum kk. Similarly, 𝒃^βq()\hat{\bm{b}}^{(\dagger)}_{\beta q} is the annihilation (creation) operator for a phonon with mode β\beta and momentum qq. The coupling between electrons and phonons is quantified by the general interband matrix elements gkqβmng^{mn}_{kq\beta}. The EPC matrix elements gkqβmng^{mn}_{kq\beta} for electron-phonon scattering between bands mm and nn, in terms of electron Bloch states read:

gkqβmn=2Mωqβumk|qH|unk+q,g^{mn}_{kq\beta}=\sqrt{\frac{\hbar}{2M\omega_{q\beta}}}\bra{u_{mk}}\partial_{q}H\ket{u_{nk+q}}, (23)

where H=iEi|ψiψi|H=\sum_{i}E_{i}\ket{\psi_{i}}\bra{\psi_{i}} is the many-body Hamiltonian of the system combining the electron and phonon degrees of freedom, |ψi\ket{\psi_{i}} are the many-body ground and excited eigenstates, and MM is the ionic effective mass. In the case of the polyacenes, the effective mass is dominated by the heavier carbon atoms.

To make our discussion concrete, we will consider a two-band model a conduction band cc and a valence band vv. This regime is applicable to the polyacene chains discussed in the main text. Correspondingly, we define gkqβcgkqβccg_{kq\beta c}\equiv g^{cc}_{kq\beta} and gkqβvgkqβvvg_{kq\beta v}\equiv g^{vv}_{kq\beta}. We further define a pair operator basis as:

𝒂^ck+q𝒂^ck=l𝑷^k+q,l𝑷^l,k,𝒂^vk+q𝒂^vk=l𝑷^l,k+q𝑷^k,l,\displaystyle\hat{\bm{a}}^{\dagger}_{ck+q}\hat{\bm{a}}_{ck}=\sum_{l}\hat{\bm{P}}^{\dagger}_{k+q,l}\hat{\bm{P}}_{l,k},\quad\hat{\bm{a}}^{\dagger}_{vk+q}\hat{\bm{a}}_{vk}=\sum_{l}\hat{\bm{P}}_{l,k+q}\hat{\bm{P}}^{\dagger}_{k,l}, (24)

and we rewrite the electron-phonon coupling in this basis as:

Hel-ph=k,l,q,β(gkqβc𝑷^k+q,l𝑷^l,k+gkqβv𝑷^l,k+q𝑷^k,l)(𝒃^β,q+𝒃^β,q).\displaystyle H_{\text{el-ph}}=\sum_{k,l,q,\beta}\left(g_{kq\beta c}\hat{\bm{P}}^{\dagger}_{k+q,l}\hat{\bm{P}}_{l,k}+g_{kq\beta v}\hat{\bm{P}}_{l,k+q}\hat{\bm{P}}^{\dagger}_{k,l}\right)\left(\hat{\bm{b}}_{\beta,q}+\hat{\bm{b}}^{\dagger}_{\beta,-q}\right). (25)

We can then rewrite the Hamiltonian in the exciton basis:

Hex-ph\displaystyle H_{\text{ex-ph}} =Q,q,β𝒟Qqβμν𝑿^Q+qμ𝑿^Qν(𝒃^β,q+𝒃^β,q),\displaystyle=\sum_{Q,q,\beta}\mathcal{D}^{\mu\nu}_{Qq\beta}\hat{\bm{X}}^{\mu\dagger}_{Q+q}\hat{\bm{X}}^{\nu}_{Q}\left(\hat{\bm{b}}_{\beta,q}+\hat{\bm{b}}^{\dagger}_{\beta,-q}\right), (26)
𝒟Qqβμν\displaystyle\mathcal{D}^{\mu\nu}_{Qq\beta} =k(gkqβcψQ+qμ(k12Q+12q)ψQν(k12Q)gkqβvψQ+qμ(k+12Q12q)ψQν(k+12Q)),\displaystyle=\sum_{k}\left(g_{kq\beta c}\psi_{Q+q\mu}\left(k-\dfrac{1}{2}Q+\dfrac{1}{2}q\right)\psi^{*}_{Q\nu}\left(k-\dfrac{1}{2}Q\right)-g_{kq\beta v}\psi_{Q+q\mu}\left(k+\dfrac{1}{2}Q-\dfrac{1}{2}q\right)\psi^{*}_{Q\nu}\left(k+\dfrac{1}{2}Q\right)\right), (27)

where the electron-phonon coupling (EPC) matrix elements gkqβvg_{kq\beta v} reflect the quantum geometry of the underlying electrons and holes Yu et al. (2024). Contributions to ExPC explicitly originate from the free-particle EPC matrix elements (gkqβvg_{kq\beta v}) and from the overlaps of excitonic envelope functions ψQ(k)\psi_{Q}(k) governed by the excitonic quantum geometry that was defined in the previous section. In the excitons considered in our work, ψQ(k)\psi_{Q}(k) is almost identical for both inverse ratios t1/t2t_{1}/t_{2} and t2/t1t_{2}/t_{1}, yet the EPC part, gkqβvg_{kq\beta v}, changes significantly. To understand this relation, we note that by considering the Hamiltonian derivatives qH\partial_{q}H within a Gaussian approximation for effective hopping parameters tij(x)t_{ij}(x) under a phonon displacement of magnitude xx, tij(x)=tijeγx2t_{ij}(x)=t_{ij}e^{-\gamma x^{2}}, following Ref. Yu et al. (2024), the geometric contributions to EPC matrix elements can be approximated as:

|gkqβvgeo|22Mωqβ(γ(μgijμν,e(𝐤)+)).|g^{\text{geo}}_{kq\beta v}|^{2}\approx\frac{\hbar}{2M\omega_{q\beta}}\Bigg{(}\gamma\Big{(}\sum_{\mu}g^{\mu\nu,e}_{ij}({\bf k})+\ldots\Big{)}\Bigg{)}. (28)

In the above, consistently with Ref. Yu et al. (2024), we recognise the presence and the significance of an electronic multiband quantum metric gijμν,e(k)=Rekiukν|1P^μ|kjukν{g^{\mu\nu,e}_{ij}(k)=\text{Re}~\bra{\partial_{k_{i}}u^{\nu}_{k}}1-\hat{P}_{\mu}\ket{\partial_{k_{j}}u^{\nu}_{k}}}, with P^μ=|ukμukμ|\hat{P}_{\mu}=\ket{u^{\mu}_{k}}\bra{u^{\mu}_{k}} a projector onto the electronic band with index μ\mu. On combining with the ExPC equation, this demonstrates the importance of quantum metric contributions to the exciton-phonon coupling, in particular contributed by the electronic quantum metric. Importantly, the electrons with the non-trivial topological invariant, will significantly contribute with the highlighted geometric terms to the enhancement of both EPC and ExPC in the topological (obstructed) electronic phase.

In the calculations for exciton dephasing, diffusion, and polaron shift, we define realistic values of γ\gamma for acoustic and optical phonons according to previous calculations/experiments on oligoacene semiconductors Thompson et al. (2023b), obtaining realistic values for the exciton linewidths. We note however that our focus is primarily on the relative difference between different transport phenomena in topological and trivial regimes rather than predicting the absolute values.

IX.6 Exciton-polaron formation

The full Hamiltonian describing a system hosting excitons and phonons can be written as

H=Hex,0+Hph,0+Hex-ph.\displaystyle H=H_{\text{ex},0}+H_{\text{ph},0}+H_{\text{ex-ph}}. (29)

To describe the impact of phonons on the excitonic properties, we define a new polaronic Hamiltonian which absorbs the impact of the exciton-phonon coupling into the single-particle energies. Following Ref. Knorr et al. (2024), we define a polaronic transformation:

S=Q,qν,β𝒟Qqβμν(1EνQ+qEμQ+ωqβ𝒃^β,q+1EνQ+qEμQωqβ𝒃^β,q)𝑿^Q+qν𝑿^Qμ,\displaystyle S=\sum_{Q,q\nu,\beta}\mathcal{D}^{\mu\nu}_{Qq\beta}\left(\dfrac{1}{E_{\nu{Q+q}}-E_{\mu Q}+\hbar\omega_{q\beta}}\hat{\bm{b}}^{\dagger}_{\beta,-q}+\dfrac{1}{E_{\nu{Q+q}}-E_{\mu Q}-\hbar\omega_{q\beta}}\hat{\bm{b}}_{\beta,q}\right)\hat{\bm{X}}^{\nu\dagger}_{Q+q}\hat{\bm{X}}^{\mu}_{Q}, (30)

which allows us to rewrite the Hamiltonian as:

H~=Hex,0+Hph,012[S,Hex-ph].\displaystyle\tilde{H}=H_{\text{ex},0}+H_{\text{ph},0}-\dfrac{1}{2}\left[S,H_{\text{ex-ph}}\right]. (31)

On solving the commutator, we arrive at the following Hamiltonian:

H~=Hex,0+Hph,0Q,qν,β|𝒟Qqβμν|2(nqβ+1EνQ+qEμQ+ωqβ+nqβEνQ+qEμQωqβ)𝑿^Qμ𝑿^Qμ.\displaystyle\tilde{H}=H_{\text{ex},0}+H_{\text{ph},0}-\sum_{Q,q\nu,\beta}|\mathcal{D}^{\mu\nu}_{Qq\beta}|^{2}\left(\dfrac{n^{\beta}_{q}+1}{E_{\nu{Q+q}}-E_{\mu Q}+\hbar\omega_{q\beta}}+\dfrac{n^{\beta}_{q}}{E_{\nu{Q+q}}-E_{\mu Q}-\hbar\omega_{q\beta}}\right)\hat{\bm{X}}^{\mu\dagger}_{Q}\hat{\bm{X}}^{\mu}_{Q}. (32)

The Hamiltonian H~\tilde{H} can be solved for the phonon-interaction corrected excitonic envelopes ψ~μQ(𝐤)\tilde{\psi}_{\mu Q}({\bf k}) on achieving self-consistency with the calculated self-energies ΣμQ\Sigma_{\mu Q}, the associated dephasing rates ΓμQ=ImΣμQ\Gamma_{\mu Q}=\text{Im}~\Sigma_{\mu Q}, and the given exciton-phonon interaction matrix elements 𝒟Qqβμν\mathcal{D}^{\mu\nu}_{Qq\beta}. Namely, we have E~μQ=EμQReΣμQ{\tilde{E}_{\mu Q}=E_{\mu Q}-\text{Re}~\Sigma_{\mu Q}}, with:

ReΣμQ=limδ00q,ν,β|𝒟Qqβμν|2(nqβ+1EνQ+qEμQ+ωqβ+iΓQ+iδ0+nqβEνQ+qEμQωqβ+iΓQ+iδ0).\text{Re}~\Sigma_{\mu Q}=-\lim_{\delta_{0}\rightarrow 0}\Re\sum_{q,\nu,\beta}|\mathcal{D}^{\mu\nu}_{Qq\beta}|^{2}\left(\dfrac{n^{\beta}_{q}+1}{E_{\nu{Q+q}}-E_{\mu Q}+\hbar\omega_{q\beta}+\text{i}\Gamma_{Q}+\text{i}\delta_{0}}+\dfrac{n^{\beta}_{q}}{E_{\nu{Q+q}}-E_{\mu Q}-\hbar\omega_{q\beta}+\text{i}\Gamma_{Q}+\text{i}\delta_{0}}\right). (33)

We observe a clear polaron shift, as shown in Fig. 5 of the main text, and a minor renormalisation of the excitonic effective mass. The excitonic mass renormalisation arises from the Feynman diagrams associated with the coupling of the virtual phonon cloud to the excitons Burovski et al. (2008). Finally, on differentiating the polaron-renormalised band energy E~μQ\tilde{E}_{\mu Q}, we obtain:

v~μQ=1QE~μQ,\langle\tilde{v}_{\mu Q}\rangle=\frac{1}{\hbar}\partial_{Q}\tilde{E}_{\mu Q}, (34)

the polaron-renormalised exciton group velocities v~μQ\langle\tilde{v}_{\mu Q}\rangle. Here, implicitly, the derivatives of the matrix elements 𝒟Qqβμν\mathcal{D}^{\mu\nu}_{Qq\beta} entering the self-energy ΣμQ\Sigma_{\mu Q} that satisfies a self-consistency condition, allow the excitonic quantum geometry to affect the renormalised exciton transport in the presence of a phonon cloud.

Having considered the effects of the virtual phonons on the exciton masses and velocities, we moreover consider an expectation value v~μQ2\langle\tilde{v}^{2}_{\mu Q}\rangle. Analogously as in the main text, this quantity enters the phonon-mediated diffusivity that accounts for a polaron shift D~ph\tilde{D}_{\text{ph}}, which is mediated by the temperature-dependent scattering of exciton-polarons from the thermally-populated phonons:

D~ph=Q,νv~νQ2ΓQeβE~νQ𝒵,\tilde{D}_{\text{ph}}=\sum_{Q,\nu}\frac{\langle\tilde{v}^{2}_{\nu Q}\rangle}{\Gamma_{Q}}\frac{e^{-\beta\tilde{E}_{\nu Q}}}{\mathcal{Z}}, (35)

with thermodynamic β=1kBT\beta=\frac{1}{k_{B}T}, and 𝒵\mathcal{Z} a partition function for exciton-polaron states. Using a derivation analogous to that in Eq. (21) for the group velocity of excitons, we find that for the polaronic states we can write:

v~μQ2=v~μQ2+12νμ|Δ~Qμν|2g~xxμν(Q),\langle\tilde{v}^{2}_{\mu Q}\rangle=\langle{\tilde{v}}_{\mu Q}\rangle^{2}+\frac{1}{\hbar^{2}}\sum_{\nu\neq\mu}|\tilde{\Delta}^{\mu\nu}_{Q}|^{2}\tilde{g}^{\mu\nu}_{xx}(Q), (36)

with Δ~Qμν=E~μQE~νQ\tilde{\Delta}^{\mu\nu}_{Q}=\tilde{E}_{\mu Q}-\tilde{E}_{\nu Q}, which allows the renormalised multiband exciton quantum metric elements g~xxμν(Q)\tilde{g}^{\mu\nu}_{xx}(Q) to modify the phonon-mediated diffusion via interband velocity matrix elements. Intuitively, the latter determine the variance of the renormalised velocity operator. On substituting the exciton quantum metric-dependent v~μQ2\langle\tilde{v}^{2}_{\mu Q}\rangle for D~ph\tilde{D}_{\text{ph}}, we observe that the geometric contribution enhances the phonon-mediated diffusion of the topological exciton-polarons.

Finally, we note that in the presence of exciton-polaron corrections Knorr et al. (2024); Dai et al. (2024b), the topology of excitons remains unaltered. Furthermore, the transport in the presence of a non-uniform electric field qualitatively overlaps with the calculation which did not involve the renormalisation with phonons. We show the corresponding results in Fig. 5 of the main text.

References

  • Wang et al. (2018) Gang Wang, Alexey Chernikov, Mikhail M Glazov, Tony F Heinz, Xavier Marie, Thierry Amand,  and Bernhard Urbaszek, “Colloquium: Excitons in atomically thin transition metal dichalcogenides,” Reviews of Modern Physics 90, 021001 (2018).
  • Perea-Causin et al. (2022) Raul Perea-Causin, Daniel Erkensten, Jamie M Fitzgerald, Joshua JP Thompson, Roberto Rosati, Samuel Brem,  and Ermin Malic, “Exciton optics, dynamics, and transport in atomically thin semiconductors,” APL Materials 10 (2022).
  • Posmyk et al. (2024) Katarzyna Posmyk, Mateusz Dyksik, Alessandro Surrente, Duncan K Maude, Natalia Zawadzka, Adam Babiński, Maciej R Molas, Watcharaphol Paritmongkol, Mirosław Maczka, William A Tisdale, et al., “Exciton fine structure in 2d perovskites: The out-of-plane excitonic state,” Advanced Optical Materials 12, 2300877 (2024).
  • Mikhnenko et al. (2015) Oleksandr V Mikhnenko, Paul WM Blom,  and Thuc-Quyen Nguyen, “Exciton diffusion in organic semiconductors,” Energy & Environmental Science 8, 1867–1888 (2015).
  • Valencia et al. (2023) Ana M Valencia, Daniel Bischof, Sebastian Anhäuser, Marc Zeplichal, Andreas Terfort, Gregor Witte,  and Caterina Cocchi, “Excitons in organic materials: revisiting old concepts with new insights,” Electronic Structure 5, 033003 (2023).
  • Chernikov et al. (2014) Alexey Chernikov, Timothy C Berkelbach, Heather M Hill, Albert Rigosi, Yilei Li, Burak Aslan, David R Reichman, Mark S Hybertsen,  and Tony F Heinz, “Exciton binding energy and nonhydrogenic Rydberg series in monolayer WS2,” Physical review letters 113, 076802 (2014).
  • Yu et al. (2015) Hongyi Yu, Xiaodong Cui, Xiaodong Xu,  and Wang Yao, “Valley excitons in two-dimensional semiconductors,” National Science Review 2, 57–70 (2015).
  • Dresselhaus et al. (2007) Mildred S Dresselhaus, Gene Dresselhaus, Riichiro Saito,  and Ado Jorio, “Exciton photophysics of carbon nanotubes,” Annu. Rev. Phys. Chem. 58, 719–747 (2007).
  • Giannini et al. (2022) Samuele Giannini, Wei-Tao Peng, Lorenzo Cupellini, Daniele Padula, Antoine Carof,  and Jochen Blumberger, “Exciton transport in molecular organic semiconductors boosted by transient quantum delocalization,” Nature Communications 13, 2755 (2022).
  • Wilson et al. (2011) Mark WB Wilson, Akshay Rao, Jenny Clark, R Sai Santosh Kumar, Daniele Brida, Giulio Cerullo,  and Richard H Friend, “Ultrafast dynamics of exciton fission in polycrystalline pentacene,” Journal of the American Chemical Society 133, 11830–11833 (2011).
  • Congreve et al. (2013) Daniel N Congreve, Jiye Lee, Nicholas J Thompson, Eric Hontz, Shane R Yost, Philip D Reusswig, Matthias E Bahlke, Sebastian Reineke, Troy Van Voorhis,  and Marc A Baldo, “External quantum efficiency above 100% in a singlet-exciton-fission–based organic photovoltaic cell,” Science 340, 334–337 (2013).
  • Xu et al. (2021) Yuwei Xu, Pei Xu, Dehua Hu,  and Yuguang Ma, “Recent progress in hot exciton materials for organic light-emitting diodes,” Chemical Society Reviews 50, 1030–1069 (2021).
  • Chowdhury et al. (2025) Rituparno Chowdhury, Marco D Preuss, Hwan-Hee Cho, Joshua JP Thompson, Samarpita Sen, Tomi Baikie, Pratyush Ghosh, Yorrick Boeije, Xian-Wei Chua, Kai-Wei Chang, et al., “Circularly polarized electroluminescence from chiral supramolecular semiconductor thin films,” Science 387, 1175–1181 (2025).
  • Shanmugaraj and John (2019) Krishnamoorthy Shanmugaraj and S Abraham John, “Water-soluble MoS2 quantum dots as effective fluorescence probe for the determination of bilirubin in human fluids,” Spectrochim. Acta A Mol. Biomol. Spectrosc. 215, 290–296 (2019).
  • Geldert et al. (2017) Alisha Geldert, Chwee Teck Lim, et al., “Paper-based MoS2 nanosheet-mediated fret aptasensor for rapid malaria diagnosis,” Sci. Rep. 7, 1–8 (2017).
  • Huang et al. (2018) Yu Li Huang, Yu Jie Zheng, Zhibo Song, Dongzhi Chi, Andrew TS Wee,  and Su Ying Quek, “The organic–2d transition metal dichalcogenide heterointerface,” Chemical Society Reviews 47, 3241–3264 (2018).
  • Sneyd et al. (2021) Alexander J Sneyd, Tomoya Fukui, David Paleček, Suryoday Prodhan, Isabella Wagner, Yifan Zhang, Jooyoung Sung, Sean M Collins, Thomas JA Slater, Zahra Andaji-Garmaroudi, et al., “Efficient energy transport in an organic semiconductor mediated by transient exciton delocalization,” Science Advances 7, eabh4232 (2021).
  • Muth et al. (2024) Dominik Muth, Sebastian Anhäuser, Daniel Bischof, Anton Krüger, Gregor Witte,  and Marina Gerhard, “Transport, trapping, triplet fusion: thermally retarded exciton migration in tetracene single crystals,” Nanoscale 16, 13471–13482 (2024).
  • Ghorab et al. (2022) Mahya Ghorab, Ali Fattah,  and Mojtaba Joodaki, “Fundamentals of organic solar cells: A review on mobility issues and measurement methods,” Optik 267, 169730 (2022).
  • Gunder et al. (2021) Darius Gunder, Ana M Valencia, Michele Guerrini, Tobias Breuer, Caterina Cocchi,  and Gregor Witte, “Polarization resolved optical excitation of charge-transfer excitons in pen: Pfp cocrystalline films: limits of nonperiodic modeling,” The Journal of Physical Chemistry Letters 12, 9899–9905 (2021).
  • Chen et al. (2023) Chen Chen, Liang Wang, Yuandong Sun, Yiwei Fu, Chuanhang Guo, Bojun Zhou, Zirui Gan, Dan Liu, Wei Li,  and Tao Wang, “Realizing an unprecedented fill factor of 82.2% in ternary organic solar cells via co-crystallization of non-fullerene acceptors,” Advanced Functional Materials 33, 2305765 (2023).
  • Bettis Homan et al. (2017) Stephanie Bettis Homan, Vinod K Sangwan, Itamar Balla, Hadallia Bergeron, Emily A Weiss,  and Mark C Hersam, “Ultrafast exciton dissociation and long-lived charge separation in a photovoltaic pentacene–MoS2 van der Waals heterojunction,” Nano letters 17, 164–169 (2017).
  • Bowman et al. (2022) Alan R Bowman, Samuel D Stranks,  and Bartomeu Monserrat, “Investigation of singlet fission–halide perovskite interfaces,” Chemistry of Materials 34, 4865–4875 (2022).
  • Thompson et al. (2023a) Joshua JP Thompson, Victoria Lumsargis, Maja Feierabend, Quichen Zhao, Kang Wang, Letian Dou, Libai Huang,  and Ermin Malic, “Interlayer exciton landscape in WS 2/tetracene heterostructures,” Nanoscale 15, 1730–1738 (2023a).
  • Qi and Zhang (2011) Xiao-Liang Qi and Shou-Cheng Zhang, “Topological insulators and superconductors,” Rev. Mod. Phys. 83, 1057–1110 (2011).
  • Hasan and Kane (2010) M. Z. Hasan and C. L. Kane, “Colloquium: Topological Insulators,” Rev. Mod. Phys. 82, 3045–3067 (2010).
  • Armitage et al. (2018) N. P. Armitage, E. J. Mele,  and Ashvin Vishwanath, “Weyl and dirac semimetals in three-dimensional solids,” Rev. Mod. Phys. 90, 015001 (2018).
  • Chang et al. (2023) Cui-Zu Chang, Chao-Xing Liu,  and Allan H. MacDonald, “Colloquium: Quantum anomalous hall effect,” Rev. Mod. Phys. 95, 011002 (2023).
  • Wu et al. (2017) Fengcheng Wu, Timothy Lovorn,  and A. H. MacDonald, “Topological exciton bands in moiré heterojunctions,” Phys. Rev. Lett. 118, 147401 (2017).
  • Kwan et al. (2021) Yves H. Kwan, Yichen Hu, Steven H. Simon,  and S. A. Parameswaran, “Exciton band topology in spontaneous quantum anomalous Hall insulators: Applications to twisted bilayer graphene,” Phys. Rev. Lett. 126, 137601 (2021).
  • Jankowski et al. (2024) Wojciech J. Jankowski, Joshua J. P. Thompson, Bartomeu Monserrat,  and Robert-Jan Slager, “Excitonic topology and quantum geometry in organic semiconductors,”  (2024), arXiv:2406.11951 [cond-mat.mes-hall] .
  • Davenport et al. (2024) Henry Davenport, Johannes Knolle,  and Frank Schindler, “Interaction-induced crystalline topology of excitons,”  (2024), arXiv:2405.19394 [cond-mat.mes-hall] .
  • Zhu et al. (2024) Jianhua Zhu, Haoxiang Chen, Ji Chen,  and Wei Wu, “One-dimensional dexter-type excitonic topological phase transition,” Phys. Rev. B 110, 085418 (2024).
  • Provost and Vallee (1980) JP Provost and G Vallee, “Riemannian structure on manifolds of quantum states,” Communications in Mathematical Physics 76, 289–301 (1980).
  • Bouhon et al. (2023) Adrien Bouhon, Abigail Timmel,  and Robert-Jan Slager, “Quantum geometry beyond projective single bands,”  (2023), arXiv:2303.02180 [cond-mat.mes-hall] .
  • Cirera et al. (2020) Borja Cirera, Ana Sánchez-Grande, Bruno de la Torre, José Santos, Shayan Edalatmanesh, Eider Rodríguez-Sánchez, Koen Lauwaet, Benjamin Mallada, Radek Zbořil, Rodolfo Miranda, et al., “Tailoring topological order and π\pi-conjugation to engineer quasi-metallic polymers,” Nature nanotechnology 15, 437–443 (2020).
  • Romanin et al. (2022) D. Romanin, M. Calandra,  and A. W. Chin, “Excitonic switching across a 𝕫2{\mathbb{z}}_{2} topological phase transition: From Mott-Wannier to Frenkel excitons in organic materials,” Phys. Rev. B 106, 155122 (2022).
  • Ahn et al. (2019) Junyeong Ahn, Sungjoon Park, Dongwook Kim, Youngkuk Kim,  and Bohm-Jung Yang, “Stiefel-Whitney classes and topological phases in band theory,” Chinese Physics B 28, 117101 (2019).
  • Bouhon et al. (2020) Adrien Bouhon, Tomáš Bzdusek,  and Robert-Jan Slager, “Geometric approach to fragile topology beyond symmetry indicators,” Phys. Rev. B 102, 115135 (2020).
  • Su et al. (1979) W. P. Su, J. R. Schrieffer,  and A. J. Heeger, “Solitons in polyacetylene,” Phys. Rev. Lett. 42, 1698–1701 (1979).
  • Su et al. (1980) W. P. Su, J. R. Schrieffer,  and A. J. Heeger, “Soliton excitations in polyacetylene,” Phys. Rev. B 22, 2099–2111 (1980).
  • Thompson et al. (2023b) Joshua JP Thompson, Dominik Muth, Sebastian Anhäuser, Daniel Bischof, Marina Gerhard, Gregor Witte,  and Ermin Malic, “Singlet-exciton optics and phonon-mediated dynamics in oligoacene semiconductor crystals,” Natural Sciences 3, e20220040 (2023b).
  • Rosati et al. (2020) Roberto Rosati, Samuel Brem, Raül Perea-Causín, Robert Schmidt, Iris Niehues, Steffen Michaelis de Vasconcellos, Rudolf Bratschitsch,  and Ermin Malic, “Strain-dependent exciton diffusion in transition metal dichalcogenides,” 2D Materials 8, 015030 (2020).
  • Ashoka et al. (2022) Arjun Ashoka, Nicolas Gauriot, Aswathy V. Girija, Nipun Sawhney, Alexander J. Sneyd, Kenji Watanabe, Takashi Taniguchi, Jooyoung Sung, Christoph Schnedermann,  and Akshay Rao, “Direct observation of ultrafast singlet exciton fission in three dimensions,” Nature Communications 13 (2022), 10.1038/s41467-022-33647-5.
  • Zhang et al. (2022) Zhilong Zhang, Jooyoung Sung, Daniel TW Toolan, Sanyang Han, Raj Pandya, Michael P Weir, James Xiao, Simon Dowland, Mengxia Liu, Anthony J Ryan, et al., “Ultrafast exciton transport at early times in quantum dot solids,” Nature Materials 21, 533–539 (2022).
  • Thompson et al. (2022) Joshua JP Thompson, Samuel Brem, Marne Verjans, Robert Schmidt, Steffen Michaelis de Vasconcellos, Rudolf Bratschitsch,  and Ermin Malic, “Anisotropic exciton diffusion in atomically-thin semiconductors,” 2D Materials 9, 025008 (2022).
  • Cohen et al. (2024) Galit Cohen, Jonah B Haber, Jeffrey B Neaton, Diana Y Qiu,  and Sivan Refaely-Abramson, “Phonon-driven femtosecond dynamics of excitons in crystalline pentacene from first principles,” Physical review letters 132, 126902 (2024).
  • Yu et al. (2024) Jiabin Yu, Christopher J. Ciccarino, Raffaello Bianco, Ion Errea, Prineha Narang,  and B. Andrei Bernevig, “Non-trivial quantum geometry and the strength of electron–phonon coupling,” Nature Physics 20, 1262–1268 (2024).
  • Markvart and Greef (2004) T Markvart and R Greef, “Polaron-exciton model of resonance energy transfer,” The Journal of chemical physics 121, 6401–6405 (2004).
  • Hansen et al. (2022) Kameron R Hansen, C Emma McClure, Daniel Powell, Hao-Chieh Hsieh, Laura Flannery, Kelsey Garden, Edwin J Miller, Daniel J King, Sami Sainio, Dennis Nordlund, et al., “Low exciton binding energies and localized exciton–polaron states in 2d tin halide perovskites,” Advanced Optical Materials 10, 2102698 (2022).
  • Baranowski et al. (2024) Michal Baranowski, Andrzej Nowok, Krzysztof Galkowski, Mateusz Dyksik, Alessandro Surrente, Duncan Maude, Marios Zacharias, George Volonakis, Samuel D Stranks, Jacky Even, et al., “Polaronic mass enhancement and polaronic excitons in metal halide perovskites,” ACS Energy Letters 9, 2696–2702 (2024).
  • Knorr et al. (2024) Willy Knorr, Samuel Brem, Giuseppe Meneghini,  and Ermin Malic, “Polaron-induced changes in moiré exciton propagation in twisted van der waals heterostructures,”  (2024), arXiv:2401.07703 [cond-mat.mes-hall] .
  • Dai et al. (2024a) Zhenbang Dai, Chao Lian, Jon Lafuente-Bartolome,  and Feliciano Giustino, “Excitonic polarons and self-trapped excitons from first-principles exciton-phonon couplings,” Phys. Rev. Lett. 132, 036902 (2024a).
  • Coehoorn et al. (2017) R Coehoorn, L Zhang, PA Bobbert,  and H Van Eersel, “Effect of polaron diffusion on exciton-polaron quenching in disordered organic semiconductors,” Physical Review B 95, 134202 (2017).
  • Hurtado Parra et al. (2022) Sebastian Hurtado Parra, Daniel B Straus, Bryan T Fichera, Natasha Iotov, Cherie R Kagan,  and James M Kikkawa, “Large exciton polaron formation in 2d hybrid perovskites via time-resolved photoluminescence,” ACS nano 16, 21259–21265 (2022).
  • Sung et al. (2019) Jooyoung Sung, Christoph Schnedermann, Limeng Ni, Aditya Sadhanala, Richard Y. S. Chen, Changsoon Cho, Lee Priest, Jong Min Lim, Hyun-Kyung Kim, Bartomeu Monserrat, Philipp Kukura,  and Akshay Rao, “Long-range ballistic propagation of carriers in methylammonium lead iodide perovskite thin films,” Nature Physics 16, 171–176 (2019).
  • Schnedermann et al. (2019) Christoph Schnedermann, Jooyoung Sung, Raj Pandya, Sachin Dev Verma, Richard YS Chen, Nicolas Gauriot, Hope M Bretscher, Philipp Kukura,  and Akshay Rao, “Ultrafast tracking of exciton and charge carrier transport in optoelectronic materials on the nanometer scale,” The Journal of Physical Chemistry Letters 10, 6727–6733 (2019).
  • Yang et al. (2016) Fujiang Yang, Ruixuan Meng, Gaiyan Zhang, Kun Gao,  and Shijie Xie, “Migration of an exciton in organic polymers driven by a nonuniform internal electric field,” Organic Electronics 30, 171–175 (2016).
  • Chaudhary et al. (2021) Swati Chaudhary, Christina Knapp,  and Gil Refael, “Anomalous exciton transport in response to a uniform in-plane electric field,” Physical Review B 103 (2021), 10.1103/physrevb.103.165119.
  • Antonius and Louie (2022) Gabriel Antonius and Steven G. Louie, “Theory of exciton-phonon coupling,” Phys. Rev. B 105, 085111 (2022).
  • Burovski et al. (2008) Evgeni Burovski, Holger Fehske,  and Andrei S. Mishchenko, “Exact treatment of exciton-polaron formation by diagrammatic monte carlo simulations,” Phys. Rev. Lett. 101, 116403 (2008).
  • Dai et al. (2024b) Zhenbang Dai, Chao Lian, Jon Lafuente-Bartolome,  and Feliciano Giustino, “Excitonic polarons and self-trapped excitons from first-principles exciton-phonon couplings,” Phys. Rev. Lett. 132, 036902 (2024b).