You are on page 1of 24

Article https://doi.org/10.

1038/s41467-023-43348-2

Gravitationally induced decoherence vs


space-time diffusion: testing the quantum
nature of gravity
Received: 12 July 2023 Jonathan Oppenheim 1 , Carlo Sparaciari 1
, Barbara Šoda 1,2,3
&
Zachary Weller-Davies1,3
Accepted: 8 November 2023

We consider two interacting systems when one is treated classically while the
Check for updates other system remains quantum. Consistent dynamics of this coupling has been
1234567890():,;
1234567890():,;

shown to exist, and explored in the context of treating space-time classically.


Here, we prove that any such hybrid dynamics necessarily results in deco-
herence of the quantum system, and a breakdown in predictability in the
classical phase space. We further prove that a trade-off between the rate of this
decoherence and the degree of diffusion induced in the classical system is a
general feature of all classical quantum dynamics; long coherence times
require strong diffusion in phase-space relative to the strength of the coupling.
Applying the trade-off relation to gravity, we find a relationship between the
strength of gravitationally-induced decoherence versus diffusion of the metric
and its conjugate momenta. This provides an experimental signature of the-
ories in which gravity is fundamentally classical. Bounds on decoherence rates
arising from current interferometry experiments, combined with precision
measurements of mass, place significant restrictions on theories where Ein-
stein’s classical theory of gravity interacts with quantum matter. We find that
part of the parameter space of such theories are already squeezed out, and
provide figures of merit which can be used in future mass measurements and
interference experiments.

When considering the dynamics of composite quantum systems, there field theory, we only know how to treat space–time classically. Likewise
are many regimes where one system can be taken to be classical and in cosmology, vacuum fluctuations are a quantum effect that we
the other quantum-mechanical. For example, in quantum thermo- believe seeds galaxy formation, while the expanding space–time
dynamics, we often have a quantum system interacting with a large they live on can only be treated classically. In addition to the need for
thermal reservoir that can be treated classically, whilst in atomic an effective theory that treats space–time in the classical limit, there
physics it is common to consider the behaviour of quantum atoms in has long been a debate about whether one should quantise gravity1–12.
the presence of classical electromagnetic fields. Things become more The prevailing view has been that a quantum–classical coupling is
complicated when one considers classical-quantum (CQ) dynamics inconsistent. Many proposals for such dynamics13,14 are not completely
where the quantum system back-reacts on the classical system. This is positive (CP), meaning they are at best an approximation and fail
particularly relevant in gravity because we would like to study the outside a regime of validity15,16. A map Λ is completely positive (CP), iff
back-reaction of thermal radiation being emitted from black holes in 1  Λ is positive. This is the required condition used to derive the GKSL
space–time, and while the matter fields can be described by quantum equation. If it is violated, the dynamics acting on half of an entangled

1
Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK. 2Department of Physics, University of Waterloo,
Waterloo, ON, Canada. 3Perimeter Institute for Theoretical Physics, Waterloo, ON, Canada. e-mail: j.oppenheim@ucl.ac.uk

Nature Communications | (2023)14:7910 1


Article https://doi.org/10.1038/s41467-023-43348-2

state, give negative probabilities. The semi-classical Einstein’s the dynamics means its interaction with quantum fields necessarily
equation17–19, which replaces the quantum operator corresponding to results in unpredictability and gravitationally induced decoherence.
the stress-energy tensor by its expectation value, is another attempt to This trade-off between the decoherence rate and diffusion pro-
treat the classical limit from an effective point of view, but it is non- vides an experimental signature, not only of models of hybrid New-
linear in the state, leading to pathological behaviour if quantum fluc- tonian dynamics such as refs. 24,33 or post-quantum theories of
tuations are of comparable magnitude to the stress-energy tensor20. general relativity such as refs. 11,39 but of any theory which treats
This is often the precise regime we would like to understand. gravity as being fundamentally classical. The metric and their con-
However, examples of classical–quantum dynamics such as those jugate momenta necessarily diffuse away from what Einstein’s general
first introduced in refs. 21,22 and studied in refs. 11,23–27 do not suffer relativity predicts. This experimental signature squeezes
from such problems and are consistent. More generally, the master classical–quantum theories of gravity from both sides: if one has
equation shown in Eq. (4)11, is linear in the state space, preserves the shorter decoherence times for superpositions of different mass dis-
division of classical degrees of freedom and quantum ones, and is tributions, one necessarily has more diffusion of the metric and con-
completely positive (CP), and preserves normalisation. This ensures jugate momenta. In the “Methods” subsection “Detecting gravitational
that probabilities of measurement outcomes remain positive and diffusion” we show that the latter effect causes imprecision in mea-
always add to 1. The dynamics is related to the GKSL or Lindblad surements of mass such as those undertaken in the Cavendish
equation28,29, which for bounded generators of the dynamics, is the experiment49–51 or in measurements of Newton’s constant “Big G"52–54.
most general Markovian dynamics for an open quantum system. The precision at which a mass can be measured in a short time, thus
More precisely, we consider dynamics which is autonomous, meaning provides an upper bound on the amount of gravitational diffusion, as
the couplings in the theory do not depend on time. Likewise, Eq. (4) is quantified by Eq. (42). In the other direction, decoherence experiments
the most general Markovian classical–quantum dynamics with boun- place a lower bound on the diffusion. Our estimates suggest that
ded generators11,30. Sub-classes of this master equation, along with experimental lower bounds on the coherence time of large
measurement and feedback approaches, have been discussed in the molecules55–60, combined with gravitational experiments measuring
context of Newtonian models of gravity24,25,31–35 and further developed the acceleration of small masses61–63, already place strong restrictions
into a spatially covariant framework so that Einstein gravity in the ADM on theories where space–time is not quantised. In the section
formalism36 emerges as a limiting case11,37. Dynamics which is mani- “Physical constraints on the classicality of gravity” we show that several
festly diffeomorphism invariant has also been introduced using path realisations of CQ-gravity are already ruled out, while other realisations
integral methods38,39. produce enough diffusion away from general relativity to be detect-
In this work, we move away from specific realisations of CQ able by future table-top experiments. Although the absence of such
dynamics, in order to discuss their common features and the experi- deviations from general relativity would not be as direct a confirmation
mental signatures that follow from this. An early precursor to the of the quantum nature of gravity, such as experiments proposed in
discussion here is the insight of Diósi22, who added classical noise and refs. 64–74 to exhibit entanglement or coherence generated by grav-
quantum decoherence to the master equation of ref. 13, and found itons, it would effectively rule out any sensible theory that treats
the noise and decoherence trade-off required for the dynamics to space–time classically. While confirmation of gravitational diffusion
become completely positive. Here we prove that the phenomena would suggest that space-time is fundamentally classical. Experiments
found in refs. 11,22,26 are generic features of all CQ dynamics; to detect or bind gravitational diffusion also provide immediate-term
the classical–quantum interaction necessarily induces decoherence on prospects for probing the quantumness of gravity, while
the quantum system, and there is a generic trade-off between the entanglement-based experiments will only be feasible in the long term.
rate of decoherence and the amount of diffusion in the classical phase The outline of the remainder of this paper is as follows. In the
space. The stronger the interaction between the quantum system and subsection “Classical-quantum dynamics” we review the general form
the classical one, the greater the trade-off. One cannot have quantum of the CQ master equation of classical–quantum systems as derived in
systems with long-coherence times without inducing a lot of diffusion refs. 11,30. The CQ-map can be represented in a manner akin to the
in the classical system. One can also generalise this result to a trade-off Kraus representation75 for quantum maps, with conditions for it to be
between the rate of diffusion and the strength of more general completely positive and trace preserving (CPTP). We can perform a
couplings to Lindblad operators, with decoherence being a short time moment expansion of the CQ-map taking states at some
special case. initial time, to states at a later time. This gives us the CQ version of the
Kramers–Moyal expansion76,77, presented in the subsection “The CQ
Results Kramers–Moyal expansion”. The physical meaning of the moments is
Our main result is expressed as Eqs. (26) and (23), which bounds given in the subsection “Physical interpretation of the moments”. Our
the product of diffusion coefficients and Lindblad coupling constants main result is presented in the section “A trade-off between deco-
in terms of the strength of the CQ-interaction. It is precisely this herence and diffusion”, where we show that there is a general trade-off
trade-off which allows the theories considered here, to evade the between decoherence of the quantum system and diffusion in the
no-go arguments of Feynmann1,2, Aharonov3, Eppley and Hannah4 classical system. We generalise the trade-off to the case of fields in the
and others1,7–9,15,16,40–48. The essence of arguments against subsection “Trade-off in the presence of fields” and in the subsection
quantum–classical interactions is that they would prohibit super- “Physical constraints on the classicality of gravity”, we apply the
positions of quantum systems that source a classical field. Since dif- inequality in the gravitational setting. The positivity constraints mean
ferent classical fields are perfectly distinguishable in principle, if that the considerations do not depend on the specifics of the theory,
the classical field is in a distinct state for each quantum state in the only that it treats gravity classically, and is time-local. This allows us to
superposition, the classical field could always be used to determine the discuss some of the observational implications of this result and we
state of the quantum system, causing it to decohere instantly. By comment on the relevant figures of merit required in interference and
satisfying the trade-off, the quantum system preserves coherence precision mass measurements in order to constrain theories of gravity,
because diffusion of the classical degrees of freedom means that the as they are not always readily available in published reports. In addition
state of the classical field does not determine the state of the quantum to table-top constraints, we consider those due to cosmological
system11,25. Equation (26) and other variants we derive, quantify the observations. We then conclude with a discussion of our results. The
amount of diffusion required to preserve any amount of coherence. If “Methods” section collects or previews a number of technical results.
space–time curvature is treated classically, then complete positivity of Since this paper appeared on the arXiv, we have found that when the

Nature Communications | (2023)14:7910 2


Article https://doi.org/10.1038/s41467-023-43348-2

decoherence-diffusion trade-off is saturated, there are two important The choice of basis Lμ is arbitrary, although there may be one
consequences. The first is that in the continuous class of dynamics, the which allows for unique trajectories26. Equation (1) can be viewed as a
quantum state remains pure, conditioned on the classical trajectory78. generalisation of the Kraus decomposition theorem.
The second is that in the path integral formulation, one can show that In the case where the classical degrees of freedom are taken to be
the dynamics are completely positive from the path integral alone39. discrete, Poulin25 used the diagonal form of this map to derive the most
For a generic path integral of Feynman–Vernon form79, one typically general form of Markovian master equation for bounded operators,
only knows that the dynamics are completely positive if it was derived which is the one introduced in ref. 21. When the classical degrees of
from a CPTP master equation. freedom are taken to live in a continuous configuration space, we need
to be a little more careful, since ϱ(z) may only be defined in a dis-
Classical–quantum dynamics tributional sense; for example, ϱðzÞ = δðz, zÞϱðzÞ. In this case (1) is
μ
Let us first review the general map and master equation governing completely positive if the eigenvalues of Λμν ðzjz 0 , δtÞ, λ ðzjz 0 , δtÞ, are
R 0 0 μ 0
classical–quantum dynamics. The classical degrees of freedom are positive so that dzdz P μ ðz, z Þλ ðzjz , δtÞ ≥ 0 for any vector with
described by a differential manifold M and we shall generically denote positive components P μ ðz, z 0 Þ30. One can derive the CQ master equa-
elements of the classical space by z. For example, we could take the tion by performing a short time expansion of Eq. (1) in the case when
classical degrees of freedom to be position and momenta in which case the Lμ is bounded11. To do so, we first introduce an arbitrary basis of
M = R2 and z = (q, p). The quantum degrees of freedom are described traceless Lindblad operators on the Hilbert space, Lμ = {I, Lα}. Now, at
by a Hilbert space H. Given the Hilbert space, we denote the set of δt = 0, we know Eq. (1) is the identity map, which tells us that
positive semi-definite operators with trace at most unity as S ≤ 1 ðHÞ. Λ00 ðzjz 0 , δt = 0Þ = δðz, z 0 Þ. Looking at the short-time expansion coeffi-
Then the CQ object defining the state of the CQ system at a given time cients, by Taylor expanding in δt ≪ 1, we can write
is a map ϱ : M ! S ≤ 1 ðHÞ subject to a normalisation constraint
R μ ν
M dzTrH ½ϱ = 1. To put it differently, we associate to each classical Λμν ðzjz 0 , δtÞ = δ 0 δ 0 δðz, z 0 Þ + W μν ðzjz 0 Þδt + Oðδt 2 Þ: ð3Þ
degree of freedom an un-normalised density operator, ϱ(z), such that
TrH ½ϱ = pðzÞ ≥ 0 is a normalised probability
R distribution over the clas- By substituting the short-time expansion coefficients into Eq. (1)
sical degrees of freedom and M dzϱðzÞ is a normalised density and taking the limit δt → 0 we can write the master equation in the form
operator on H. An example of such a CQ-state is the CQ qubit depicted
Z
as a 2 × 2 matrix over phase space26. More generally, we can define any ∂ϱðz, tÞ 1
= dz 0 W μν ðzjz 0 ÞLμ ϱðz 0 ÞLyν  W μν ðzÞfLyν Lμ , ϱg + , ð4Þ
CQ operator f(z) which lives in the fibre bundle with base space M and ∂t 2
fibre H.
Just as the Lindblad equation is the most general evolution law where {, }+ is the anti-commutator, and preservation of normalisation
that maps density matrices to density matrices, we can ask what is the under the trace and ∫ dz defines
most general evolution law that preserves the quantum-classical state- Z
space. Any such dynamics, if it is to preserve probabilities, must be
W μν ðzÞ = dz 0 W μν ðz 0 jzÞ: ð5Þ
completely positive, norm-preserving, and linear in the CQ-state. That
dynamics must be linear can be seen as follows: if someone prepares a
system in one of two states σ0 or σ1 depending on the value of a coin We see the CQ master equation is a natural generalisation of the
toss (j0ih0j with probability p,j1ih1j with probability 1−p), then the Lindblad equation and classical rate equation in the case of
evolution L of the system must satisfy pj0ih0j  Lσ 0 + ð1  pÞj1ih1j  classical–quantum coupling. We give a more precise interpretation of
Lσ 1 = Lð pσ 0 + ð1  pÞσ 1 Þ otherwise the system evolves differently the different terms arising when we perform the Kramers–Moyal
depending on whether we are aware of the value of the coin toss. A expansion of the master equation at the end of the section. The posi-
violation of linearity further implies that when the system is in state σ0 tivity conditions from Eq. (1) transfer to positivity conditions on the
it evolves differently depending on what state the system would have master equation via (3). We can write the positivity conditions in an
been prepared in, had the coin been j1ih1j instead of j0ih0j. This illuminating form by writing the short time expansion of the transition
motivates our restriction to linear theories. We will also require the amplitude Λμν ðzjz 0 , δtÞ, as defined by Eq. (3), in block form
map to be Markovian on the combined classical–quantum system, " #
which is equivalent to requiring that there is no hidden system that μν 0 δðz, z 0 Þ + δtW 00 ðzjz 0 Þ δtW 0β ðzjz 0 Þ
acts as a memory. This is natural if the interaction is taken to be fun- Λ ðzjz , δtÞ = + Oðδt 2 Þ ð6Þ
δtW α0 ðzjz 0 Þ δtW αβ ðzjz 0 Þ
damental, but is the assumption that one might want to remove if one
thinks of the hybrid theory as an effective description. We thus take and the dynamics will be positive if and only if Λμν ðzjz 0 , δtÞ is a positive
these as the minimal requirements that any fundamental matrix. It is possible to introduce an arbitrary set of Lindblad operators
classical–quantum theory must satisfy if it is to be consistent.  and appropriately redefine the couplings W μν ðzjz 0 Þ in Eq. (4)11. For

The most general CQ-dynamics, which maps CQ states onto most purposes, we shall work with a set of Lindblad operators that
themselves can be written in the form11 includes the identity Lμ = (I, Lα); this is sufficient since any CQ master
Z equation is completely positive if and only if it can be brought to the
X form in Eq. (4), where the matrix (6) is positive.
ϱðz, t + δtÞ = dz 0 Λμν ðzjz 0 , δtÞLμ ϱðz 0 , tÞLyν ð1Þ
μν
The CQ Kramers–Moyal expansion
where the Lμ is an orthogonal basis of operators and Λμν ðzjz 0 , δtÞ is In order to study the positivity conditions it is first useful to perform a
positive semi-definite for each z, z 0 . Henceforth, we will adopt moment expansion of the dynamics in a classical-quantum version of
the Einstein summation convention so that we can drop ∑μν with the the Kramers–Moyal expansion as in ref. 11. In classical Markovian
understanding that equal upper and lower indices are presumed to be dynamics, the Kramers–Moyal expansion relates the master equation
summed over. The normalisation of probabilities requires to the moments of the probability transition amplitude and proves to
Z be useful for a multitude of reasons. Firstly, the moments are related to
observable quantities; for example, the first and second moments of
dzΛμν ðzjz 0 , δtÞLyν Lμ = I: ð2Þ
the probability transition amplitude characterise the amount of drift

Nature Communications | (2023)14:7910 3


Article https://doi.org/10.1038/s41467-023-43348-2

and diffusion in the system. This is reviewed in the subsection “Physical determines the diffusion of the classical degrees of freedom. For this
interpretation of the moments”. Secondly, the positivity conditions on discussion, we shall take the classical degrees of freedom to live in a
the master equation transfer naturally to positivity conditions on the phase space Γ = ðM, ωÞ, where ω is the symplectic form.
moments, which we can then relate to observable quantities. In the Consider R the expectation value of any CQ operator
classical–quantum case, we shall perform a short time moment OðzÞ, hOðzÞi :¼ dzTr½OðzÞϱ which does not have an explicit time
expansion of the transition amplitude Λμν ðzjz 0 , δtÞ and then show that dependence. Its evolution law can be determined via Eq. (9)
the master equation can be written in terms of these moments. We   Z
Z
then relate the moments to observational quantities, such as the dhOi ∂ϱ
= dzTr OðzÞ = dzTrϱ
decoherence of the quantum system and the diffusion in the classical dt ∂t

system. 1 αβ
i½OðzÞ, HðzÞ + Dαβ y y
0 ðzÞLβ OðzÞLα  D0 ðzÞfLα Lβ , OðzÞg + ð10Þ
We work with the form of the dynamics in Eq. (4), using an arbi- 2
trary orthogonal basis of Lindblad operators Lμ = fI, Lα g. We take the ! #
X1
∂n  
classical degrees of freedom M to be d dimensional, z = (z1, …zd), and + Dμν
n, i1 ...in ðzÞ L y
L
ν μ Oðz, tÞ ,
∂z i1 . . . ∂z in
we label the components as zi, i ∈ {1, …d}. We begin by introducing the n=1

moments of the transition amplitude W μν ðzjz 0 Þ appearing in the CQ


master Eq. (3) where we have used cyclicity of trace and integration by parts, to bring
the equation of motion into a form that would enable us to write a CQ
Z
1 version of the Heisenberg representation11 for a CQ operator. If we are
Dn,μνi ...i ðz 0 Þ:¼ dzW μν ðzjz 0 Þðz  z 0 Þi1 . . . ðz  z 0 Þin : ð7Þ
1 n n! interested in the expectation value of phase space variables OðzÞ = z i I
then Eq. (10) gives
The subscripts ij ∈ {1, …d } label the different components of the
Z h i
vectors ðz  z 0 Þ. For example, in the case where d = 2 and the classical dhz i i
= dzDμν y
1, i Tr Lν Lμ ϱðz, tÞ
ð11Þ
degrees of freedom are position and momenta of a particle, dt
z = (z1, z2) = (q, p), then we have
P
ðz  z 0 Þ = ðz 1  z 01 , z 2  z 02 Þ = ðq  q0 , p  p0 Þ. The components are then with all higher order terms vanishing, and we see that μν≠00 Dμν y
1, i hLν Lμ i
given by ðz  z 0 Þ1 = ðq  q0 Þ and ðz  z 0 Þ2 = ðp  p0 Þ. M μν 0
n, i1 ...in ðz , δtÞ is governs the average rate at which the quantum system moves the
seen to be an nth rank tensor with dn components. classical system through phase space, and with the back-reaction is
br αμ
In terms of the components Dn,μνi ...i the short time expansion of quantified by the Hermitian matrix Dαμ 1 :¼ ðD1 Þ . The force of this
1 n
the transition amplitude Λμν ðzjz 0 Þ is given by30 back-reaction is especially apparent if the equations of motion are
!
Hamiltonian in the classical limit as in ref. 11. I.e. if we define
αβ αβ
X
1
∂n H I ðzÞ :¼ h Lyβ Lα and take Dαβ j
1, i = ωi d j h with ω the symplectic form
Λμν ðzjz 0 , δtÞ = δ μ0 δ ν0 δðz, z 0 Þ + δt Dμν 0
n, i ...i ðz Þ δðz, z 0 Þ + Oðδt 2 Þ,
n=0
1 n ∂z 0i . . . ∂z 0i and dj the exterior derivative. Then Eq. (11) is analogous to Hamilton’s
1 n

ð8Þ equations, and the CQ evolution equation after tracing out the
quantum system has the form of a Liouville’s equation to first order
and the master equation takes the form11 and in the classical limit,
!
∂ϱðz, tÞ X 1
∂n   ∂ρðz, tÞ  
ð1Þ n
D00 = fH c , ρðz, tÞg + tr fH I ðzÞ, ϱðzÞg + . . . ð12Þ
= n, i1 ...in ðzÞϱðz, tÞ ∂t
∂t n=1
∂z i1 . . . ∂z in
1 αβ
 i½HðzÞ, ϱðzÞ + Dαβ y y
0 ðzÞLα ϱðzÞLβ  D0 ðzÞfLβ Lα , ϱðzÞg +
ð9Þ with ρðzÞ :¼ Tr½ϱðzÞ.
! 2
  The significance of the second moment is also seen via Eq. (10) to
X X 1
∂ n
+ ð1Þn Dμν y
n, i1 ...in ðzÞLμ ϱðz, tÞLν ,
be related to the variance of phase space variables
μν≠00 n = 1
∂z i1 . . ∂z in
. σ z i z i :¼ hz i1 z i2 i  hz i1 ihz i2 i
1 2

μ0 y
where we define the Hermitian operator HðzÞ =  2 ðD0 Lμ
i
D0μ
0 Lμ Þ dσ 2z i , zi
(which is Hermitian since Dμ0 = D 0μ*
). We see the first line of Eq. (9)
1 2
= 2hDαβ y αβ y αβ y
2, i , i Lβ Lα i + hz 2 D1, z i Lβ Lα i  hz i2 ihD1, z i Lβ Lα i
0 0 dt 1 2 1 1 ð13Þ
describes purely classical dynamics, and is fully described by the
+ hz i1 Dαβ Lyβ Lα i  hz i1 ihDαβ y
1, z i Lβ Lα i:
moments of the identity component of the dynamics Λ00 ðzjz 0 Þ. The 1, z i
2 2

second line describes pure quantum Lindbladian evolution described


by the zeroth moments of the components Λα0 ðzjz 0 Þ, Λαβ ðzjz 0 Þ; speci- In the case when D1, z i is uncorrelated with z i2 and D1, z i uncor-
fically the (block) off diagonals, Dα0
1 2
0 ðzÞ, describe the pure Hamiltonian related with z i1 , then the growth of the variance only depends on the
evolution, whilst the components Dαβ 0 ðzÞ describe the dissipative part diffusion coefficient.
of the pure quantum evolution. Note that the Hamiltonian and Lind- The zeroth moment Dαβ 0 is just the pure Lindbladian couplings.
blad couplings can depend on the classical degrees of freedom so the The simplest example is the case of a pure decoherence process with a
second line describes the action of the classical system on the quan- single Hermitian Lindblad operator L and decoherence coupling D0.
tum one. The third line contains the non-trivial classical-quantum back- Then we can define a basis fjaig via the eigenvectors of L and
reaction, where changes in the distribution over phase space are  
induced and can be accompanied by changes in the quantum state. ∂ϱ 1
a b =  i aj ½HðzÞ, ϱ jb  D0 ðLðaÞ  LðbÞÞ hajϱjbi,
2
ð14Þ
∂t 2
Physical interpretation of the moments
Let us now briefly review the physical interpretation of the moments and we see that the matrix
 elements of ϱ which quantify coherence
that will appear in our trade-off relation. In particular, the zeroth between the states jai, b decay exponentially fast with a decay rate of
moment determines the rate of decoherence (and Lindbladian cou- D0(L(a)−L(b))2. For a damping/pumping process of a quantum har-
pling more generally), the first moment gives the force exerted by the monic oscillator with Hamiltonian H = ωa†a, L↓ = a, L↑ = a†, a the crea-
quantum system on the classical system, and the second moment tion operator, and D"" ##
0 , D0 the non-zero couplings, then standard

Nature Communications | (2023)14:7910 4


Article https://doi.org/10.1038/s41467-023-43348-2

calculations26,80 show that an initial superposition p1ffiffi2 jn + mi with n, m quantum system complete positivity demands that jjO  jj > 0, meaning
2
large and n ≫ m will initially decohere at a rate of approximately the Cauchy-Schwartz inequality in Equation (18) must hold. To reach
ðD"" ##
0 + D0 Þðm + nÞ=2, and the state will eventually thermalise to a this conclusion one can insert the CQ state into the CQ
temperature of ω= logðD## ""
0 =D0 Þ. So in this case, the Lindblad Cauchy–Schwartz inequality and repeat the proof of the Pawula
couplings not only determine the rate of decoherence but also the theorem30, which must now hold once averaged over the state. By
rate at which energy is pumped into the harmonic oscillator. In the using the short-time moment expansion of Λμν ðzjz 0 Þ defined in Eq. (8)
next section, we will derive the trade-off between Lindblad couplings and using integration by parts, we then arrive at the observational
and the diffusion coefficients. Although we will sometimes refer to this trade-off between decoherence and diffusion
as a trade-off between decoherence and diffusion, this terminology
is only strictly appropriate for pure decoherence processes, while Z h iZ h i Z h i2
dzTr 2b Dμν
i*
b
j
L ϱðzÞL y
dzTr D αβ
L ϱðzÞL y
≥  dzTr bi Dμα L ϱðzÞLy  ,
more generally, it is a trade-off between Lindblad couplings and 2, ij μ ν 0 α β  1, i μ α 

diffusion coefficients. ð19Þ

A trade-off between decoherence and diffusion which must hold for any positive CQ state ϱ(z). Stripping out the bi
In this section, we present our main result by using positivity condi- vectors, (19) is equivalent to the matrix positivity condition
tions to prove the trade-off between decoherence and diffusion seen in
models such as those of refs. 11,22,26 are in fact a general feature of all y
0  2hD2 ihD0 i  hDbr br
1 ihD1 i , 8ϱðzÞ, ð20Þ
classical-quantum interactions. We shall also generalise this, and
derive a trade-off between diffusion and arbitrary Lindbladian cou-
where we define
pling strengths. The trade-off is in relation to the strength of the
dynamics and is captured by Eqs. (20), (23) and (26). In the subsection Z h i Z h i
“Trade off in the presence of fields” we extend the trade-off to the case hD0 i = dzTr Dαβ y br
0 Lα ϱðzÞLβ , hD1 ii = dzTr Dμα y
1, i Lμ ϱðzÞLα ,
where the classical and quantum degrees of freedom can be fields and Z h i ð21Þ
use this to show that treating the metric as being classical necessarily hD2 iij = dzTr Dμν y
2, ij Lμ ϱðzÞLν :
results in diffusion of the gravitational field.
There are two separate possible sources for the force (or drift) of
Since Eq. (20) holds for all states, the tightest bound is provided
the back-reaction of the quantum system on phase space—it can be
by the infimum over all states
sourced by either the D0α 1, i components or the Lindbladian components
Dαβ
1, i . We shall deal with both sources simultaneously by considering a y
CQ Cauchy-Schwartz inequality which arises from the positivity of 0  inf f2hD2 ihD0 i  hDbr br
1 ihD1 i g: ð22Þ
ϱðzÞ
Z 
Tr dzdz 0 Λμν ðzjz 0 ÞOμ ðz, z 0 Þρðz 0 ÞOyν ðz, z 0 Þ ≥ 0 ð15Þ The quantities 〈D2〉 and 〈D0〉 appearing in Eq. (20) are related to
observational quantities. In particular 〈D2〉 is the expectation value of
for any vector of CQ operators Oμ. One can verify that this must be the amount of classical diffusion which is observed and 〈D0〉 is related
positive directly from the positivity conditions on Λμν ðzjz 0 Þ and we go to the amount of decoherence on the quantum system. The expecta-
through the details in the Appendix section “Positivity conditions and tion value of the back-reaction matrix hDbr 1 i quantifies the amount of
the trade-off between decoherence and diffusion”. A common choice back-reaction on the classical system. In the trivial case Dbr
1 = 0, Eq. (20)
for Oμ would be the set of operators Lμ = fI, Lα g appearing in the master places little restriction on the diffusion and Lindbladian rates appear-
equation. ing on the left-hand side. We already knew from refs. 28,29 that the Dαβ 0
The inequality in Eq. (15) turns out to be especially useful since it must be a positive semi-definite matrix, and we also know that diffu-
can be used to define a (pseudo) inner product on a vector of opera- sion coefficients must be positive semi-definite. However, in the non-
tors with components Oμ via trivial case, the larger the back-reaction exerted by the quantum sys-
tem, the stronger the trade-off between the diffusion coefficients and
Z h i
 i=
 ,O Lindbladian coupling. Equation (20) gives a general trade-off between
hO dzdz 0 Tr Λμν ðzjz 0 ÞO1μ ϱðz 0 ÞOy2ν ð16Þ
1 2 observed diffusion and Lindbladian rates, but we can also find a trade-
off in terms of a theory’s coupling coefficients alone. We show in the
pffiffiffiffiffiffiffiffiffiffiffiffiffi
 = hO,
where jjOjj  Oi  ≥ 0 due to Eq. (15). Technically this is not positive Appendix section “General trade-off between decoherence and diffu-
definite, but this shall not be important for our purpose. Taking the sion coefficients” that the general matrix trade-off
combination Oμ = jjO  jj2 O  hO
 ,O
 iO for vectors O1μ, O2μ, posi-
2 1μ 1 2 2μ
1 bry
tivity of the norm gives Dbr
1 D0 D1  2D2 ð23Þ
 
 2 = jj jjO
jjOjj   hO
 jj2 O  ,O iO jj2 holds for the matrix whose elements are the couplings Dμν αμ αβ
2 1 1 2 2 2, ij , D1, i , D0
  ð17Þ for any CQ dynamics. Moreover, ðI  D0 D1 ÞD br
 jj2 jjO
= jjO  jj2 jjO
 jj2  jhO
 ,O  ij2 ≥ 0, 0 1 = 0, which tells us that
2 1 2 1 2
D0 cannot vanish if there is non-zero back-reaction. Equation (23)
 jj ≠ 0 we have a Cauchy–Schwartz inequality
and as long as jjO quantifies the required amount of decoherence and diffusion in order
2
for the dynamics to be completely positive. In Eq. (23), and
 jj2 jjO
jjO  jj2  jhO
 ,O
 ij2 ≥ 0: ð18Þ throughout, D1 αβ αβ
1 2 1 2 0 is the generalised inverse of D0 , since D0 is only
required to be positive semi-definite. In the special case of a single
We can use Eq. (18) to get a trade-off between the observed dif- Lindblad operator α = 1 and classical degree of freedom, and when the
α i
fusion and decoherence by picking O2μ = δ μ Lα and O1μ = b ðz  z 0 Þi Lμ , only non-zero couplings are D11 00 0
0 :¼ D0 , D2, pp :¼ 2D2 and D1, q = 1 this
where Lμ = fI, Lα g are the Lindblad operators appearing in the master trade-off reduces to the condition D2D0 ≥ 1 used in ref. 22.
i
equationR and bh are the components
i of an arbitrary vector. In this case As a more general example, let us consider the class of theories
 jj = dzTr Dαβ L ϱLy and one can verify using CQ Pawula
jjO that are continuous in phase space, and whose back-reaction is gen-
2 0 α β
theorem30 that in order to have non-trivial back-reaction on the erated by a classical-quantum Hamiltonian H ^ ðmÞ which is only a

Nature Communications | (2023)14:7910 5


Article https://doi.org/10.1038/s41467-023-43348-2

function of the canonical coordinates qi30. These are given by classicality of the two systems can be written11
Z
∂ϱ h ðmÞ i 1 n ðmÞ o 1 n o
^ ,ϱ +
= fH c , ϱg  i H H^ ,ϱ  ^ ðmÞ
ϱ, H ρðz, tÞ = dz 0 dxdyΛμν ðzjz 0 , t; x, yÞLμ ðx, z, z 0 Þϱðz 0 , 0ÞLyν ðy, z, z 0 Þ, ð28Þ
∂t 2 2
" ðmÞ " ## ð24Þ
1 ∂ ^
H ^ ðmÞ
∂H
i j
+ fp , fp , D2, ij ϱgg + D0, ij , ϱ, where, as is usually the case with fields, in Eq. (28) it should be impli-
2 ∂qi ∂qj
citly understood that a smearing procedure has been implemented.
We elaborate on some of the details when fields are introduced in the
where Hc is the purely classical Hamiltonian, pj are the conjugate
Appendix section “Classical-quantum dynamics with fields”. The con-
momenta, and D0 and D2 are qi dependent matrices with elements D0,ij
dition for Eq. (28) to be completely positive on all CQ states is that for
and D2,ij. Then the trade-off (23) implies that they must obey the matrix
all vectors at x with components Aν ðy, z, z 0 Þ
equation 2D2D0 ⪯ 1.
It is also useful to try to obtain an observational trade-off in terms Z
of the total drift due to back-reaction as calculated in Eq. (11) dzdxdyA*μ ðx, z, z 0 ÞΛμν ðzjz 0 ; x, yÞAν ðy, z, z 0 Þ ≥ 0 ð29Þ

X Z h i
hDT1 ii = dzTr Dμν y meaning that Λμν(x, y) can be viewed as a positive matrix in μν and a
1, i Lμ ϱðzÞLν : ð25Þ
μν≠00 positive kernel in x, y. In the field-theoretic case, one can still perform
a Kramers–Moyal expansion and find a trade-off between the
It follows directly from Eq. (20) that when the back-reaction is coefficients D0(x, y), D1(x, y), D2(x, y) appearing in the master equation.
αβ
sourced by either D0μ1, i or D1, i we can arrive at the observational trade- The coefficients now have an x, y dependence, due to the spatial
off in terms of the total drift dependence of the Lindblad operators. The coefficients
D1(x, y), D2(x, y) still have a natural interpretation as measuring the
y
ð26Þ amount of force (drift) and diffusion, whilst D0(x, y) describes the
0  8hD2 ihD0 i  hDT1 ihDT1 i , 8ϱðzÞ,
purely quantum evolution on the system and can be related to
decoherence.
where the quantities appearing in Eq. (26) are now all observational
Using the positivity condition in Eq. (29) we find the same trade of
quantities, related to drift, decoherence and diffusion as outlined in
between coupling constants in Eq. (23) but where now D2(x, y) is the
the previous subsection “Physical interpretation of the moments”. We
(p + 1)n × (p + 1)n matrix-kernel with elements Dμν br
2, ij ðx, yÞ, D1 ðx, yÞ is the
believe that Eq. (26) should hold more generally, though we don’t have
(p + 1)n × p matrix-kernel with rows labelled by μi, columns labelled by
a general proof.
β, and elements Dμβ 1, i ðx, yÞ, and D0(x, y) is the p × p decoherence matrix-
In the case where the back-reaction is Hamiltonian at first order in
kernel with elements Dαβ 0 ðx, yÞ. Here i ∈ {1, …, n}α ∈ {1, …, p} and
the sense of Eq. (12), then Eq. (26) can be written as
μ ∈ {1, …, p + 1}. In the field-theoretic trade off we are treating the
objects in Eq. (23) as matrix-kernels, so that for any position-
y μ R
∂H ∂H dependent vector bμ ðxÞ, ðD2 bÞi ðxÞ = dyDμν
i j
ω  !I ω  !I  8hD2 ihD0 i, 8ϱðzÞ: ð27Þ 2, ij ðx, yÞbR ν ðyÞ, whilst for any
∂z ∂z position-dependent vector aβ ðxÞ, ðD0 aÞα ðxÞ = dyDαβ 0 ðx, yÞaβ ðyÞ.
Explicitly, we find that positivity of the dynamics is equivalent to the
As a result, we can derive a trade-off between diffusion and matrix condition
decoherence for any theory that reproduces this classical limit and " #
Z 
treats one of the systems classically. * * 2D2 ðx, yÞ Dbr
1 ðx, yÞ bðyÞ
To summarise, whenever the back-reaction of the quantum sys- dxdy½b ðxÞ, a ðxÞ ≥0 ð30Þ
Dbr
1 ðx, yÞ D0 ðx, yÞ aðyÞ
tem on the classical system induces a force on the phase space, then we
i
have a trade-off between the amount of diffusion on the classical which should be positive for any position-dependent vectors bμ ðxÞ and
system and the strength of decoherence on the quantum system (or aα(x). This is equivalent to trade-off between coupling constants in
more precisely the strength of the Lindbladian couplings Dαβ 0 ). This can Eq. (23) if we view (23) as a matrix-kernel equation.
be expressed both as a condition on the matrix of coupling co- In order for the theory to be diffeomorphism invariant, we expect
efficients in the master equation, via Eq. (23) or in terms of observable D0(x, y) and D2(x, y) to approach delta functions. We will not assume
quantities using Eqs. (20) and (26). In the case when the back-reaction this, but we shall assume that the drift back-reaction is local, so that
is Hamiltonian, we further have Equation (27). We would like to apply Dbr br
1 ðx, yÞ = δðx, yÞD1 ðxÞ. As we shall see in the next section, this is a
this trade-off to the case of gravity in the non-relativistic, Newtonian natural assumption if we want to have back-reaction which is given by a
limit. In order to do so, we will need to generalise the trade-off to the local Hamiltonian. However, one might not want to assume that the
case of quantum fields interacting with classical ones, which we do in form of the Hamiltonian remains unchanged to arbitrarily small dis-
the subsection “Trade-off in the presence of fields”. The goal will be to tances. With this locality assumption, Eq. (30) gives rise to the same
understand the implications of treating the metric (or Newtonian trade-off of Eq. (23), where the trade-off is to be interpreted as a matrix
potential) as being classical by using the trade-off when the quantum kernel inequality. Writing this out explicitly we have
back-reaction induces a force on the gravitational field which, on Z Z
expectation, is the same as the weak field limit of General Relativity. dxdyai*ν ðxÞDμα 1 βν 0 i 0
dxdy2ai*μ ðxÞDμν j
1, i ðxÞðD0 Þαβ ðx, yÞD1, j ðx Þaν ðx Þ ≤ 2, ij ðx, yÞaν ðyÞ,

Trade-off in the presence of fields ð31Þ


We would like to explore the trade-off in the gravitational setting and
explore the consequences of treating the gravitational field as being where asking that this inequality holds for all vectors aiμ ðxÞ is equiva-
classical and matter quantum. Since gravity is a field theory, we must lent to the matrix-kernel trade-off condition of Eq. (23).
first discuss classical-quantum master equations in the presence of We give two examples of master equations satisfying the coupling
fields. In the field-theoretic case, both the Lindblad operators and the constant trade-off in the Appendix section “Examples of Kernels
phase space degrees of freedom can have spatial dependence, saturating the decoherence diffusion coupling constants trade-off”.
z(x), Lμ(x) and a general bounded CP map which preserves the The decoherence-diffusion trade-off tells us how much diffusion and

Nature Communications | (2023)14:7910 6


Article https://doi.org/10.1038/s41467-023-43348-2

stochasticity are required to maintain coherence when the quantum observational quantities. This trade-off puts tight observational con-
system back-reacts on the classical one. If the interaction between the straints on classical theories of gravity which we now discuss.
classical and quantum degrees of freedom is dictated by unbounded
operators, such as the mass density, then there can exist states for Physical constraints on the classicality of gravity
which the back-reaction can be made arbitrarily large. This is the case In this section, we apply the trade-off of Eq. (30) to the case of gravity.
for a quantum particle interacting with its Newtonian potential A number of classical-quantum models of Newtonian gravity have
through its mass density at arbitrarily short distances. Hence, if one been proposed24,31–33, but since the trade-offs derived in the previous
considers a particle in a superposition of two peaked mass densities, section depend only on the back-reaction, or drift term, they are
then there can be an arbitrarily large response in the Newtonian insensitive to the particulars of the theory. We shall consider the
potential around those points, and either there must be an arbitrary Newtonian, non-relativistic limit of a classical gravitational field which
amount of diffusion, or the decoherence must occur arbitrarily fast. we reproduce in the “Methods” subsection “Newtonian limit of CQ
The former is unphysical, while the latter turns out to be the case in theory”. A fuller discussion, including a derivation of the Newtonian
simple examples of theories such as those discussed in the Methods limit starting from the covariant theories of refs. 11,39 can be found in
subsection “Decoherence rates”. ref. 81. It is in taking this limit where some care should be taken, since
Since our goal is to experimentally constrain classical-quantum one is gauge fixing the full general relativistic theory. We denote Φ to
theories of gravity, we shall hereby ask that the map (28) is CP when be the Newtonian potential and in the weak field limit of General
acting on all physical states ρ. If one allows for arbitrarily peaked mass Relativity, it has a conjugate momenta we denote by πΦ. We assume:
distributions then the coupling constant trade-off of Eq. (31) should be (i) The theory satisfies the assumptions used to derive the master
satisfied. In the field-theoretic case, we can similarly find an observa- equation as in subsections “Trade-off in the presence of fields”; in
tional trade-off, relating the expected value of the diffusion matrix particular that the theory be a completely positive norm-
〈D2(x, y)〉 to the expected value of the drift in a physical state ϱ as we preserving Markovian map, and that we can perform a short-
did in subsection “A trade-off between decoherence and diffusion”. time Kramers–Moyal expansion as in “Methods” subsection
This is done explicitly in the “Methods” subsection “Classical-quantum “Classical-quantum dynamics with fields”.
dynamics with fields”, using a field-theoretic version of the (ii) We apply the theory to the weak field limit of General Relativity,
Cauchy–Schwartz inequality given by Eq.n (73), we find whereas recalled in “Methods” subsection “Newtonian limit of CQ
theory” the Newtonian potential interacts with matter through its
Z
y mass density m(x),
2hD2 ðx, xÞi dx 0 dy0 hD0 ðx 0 , y0 ÞikhDbr br
1 ðxÞihD1 ðxÞi , ð32Þ

Z
where Eq. (34) is to be understood as a matrix inequality with entries 3
H I ðΦÞ = d xΦðxÞmðxÞ: ð36Þ
Z h i
hD0 ðx, yÞi = dzTr Dαβ y
0 Lα ðxÞϱLβ ðyÞ , ð33aÞ

and the conjugate momentum to Φ satisfies


Z h i
hDbr
1 ðx, yÞii = dzTr Dμα y
1, i Lμ ðxÞϱLα ðxÞ , ð33bÞ
∇2 Φ
π_ Φ = 4πG  mðxÞ, ð37Þ
Z h i
hD2 ðx, yÞiij = dzTr Dμν y
2, ij Lα ðxÞϱLβ ðyÞ : ð33cÞ

αβ
where in the c → ∞ limit the momentum constraint πΦ ≈ 0 is imposed
Similarly, when the back-reaction is sourced by either D0μ
1, i or D1, i it and we recover Poisson’s equation for the Newtonian potential. We
follows from Eq. (32) we can arrive at the observational trade-off in assume this limit of General Relativity is satisfied on expectation, at
terms of the total drift due to back-reaction least to leading order. This may be an overly strong assumption, since
Z the weak field limit may cease to be valid at short distances when the
y
8hD2 ðx, xÞi dx 0 dy0 hD0 ðx 0 , y0 ÞikhDT1 ðxÞihDT1 ðxÞi , ð34Þ diffusion becomes large. A relativistic treatment is initiated in (J.
Oppenheim and A. Russo, manuscript in preparation). It is also worth
noting that General Relativity has not been tested at distances shorter
where than the millimeter scale, and here we assume it holds to arbitrarily
Z short distances.
h i
hDT1 ðxÞii = dzTr D0α y α0 αβ y (iii) In relating D0 to the decoherence rate of a particle in super-
1, i ðxÞϱLα ðxÞ + D1, i ðxÞLα ϱðxÞ + D1, i ðxÞLα ðxÞϱLβ ðxÞ :
position, we shall assume that the state of interest is well
ð35Þ approximated by a state living in a Hilbert space of fixed particle
number. We believe this is a mild assumption: ordinary non-
We shall now use the trade-off to study the consequences of relativistic quantum mechanics is described via a single particle
treating the gravitational field classically. We will consider the back- Hilbert space, and we frequently place composite massive
reaction of the mass on the gravitational field to be governed by the particles in superposition and they do not typically decay into
Newtonian interaction (or more accurately, a weak field limit of Gen- multiple particles.
eral Relativity). We shall then find that experimental bounds on (iv) We will assume that the diffusion kernel D2 ðΦ, x, x 0 Þ does not
coherence lifetimes for particles in superposition require large diffu- depend on πΦ i.e. it is minimally coupled. This is reasonable, since
sion in the gravitational field in order to be maintained and this can be in the purely classical case matter couples to the Newtonian
upper bounded by gravitational experiments. potential and not its conjugate momenta.
To summarise this section, we have derived the trade-off between
decoherence and diffusion for classical–quantum field theories, both With these assumptions, and treating the matter density as a
in terms of coupling constants of the theory and in terms of ^
quantum operator mðxÞ, this tells us that in order for the back-reaction

Nature Communications | (2023)14:7910 7


Article https://doi.org/10.1038/s41467-023-43348-2

term to reproduce the Newtonian interaction on average In30 it was shown that there are two classes of CQ dynamics, at
least in the sense that there are those with continuous trajectories in
Z 
  δϱ phase space and those which contain discrete jumps. For the class of
Tr fH I , ϱg = Tr
3
^
d x mðxÞ
δπ Φ continuous CQ models (see ref. 24 and Appendix section “Continuous
X Z  ð38Þ master equation”), we know that J(x, t) should be described by a white
δϱ y
d xDμν
3
= Tr ðΦ, π Φ , xÞL μ ðxÞ L ðxÞ , noise process in time, and its statistics should be independent of the
μν≠00
1, π Φ
δπ Φ ν
mass density of the particle.
then we must pick For the discrete class (see ref. 11 and J. Oppenheim, “The con-
straints of a continuous realisation of post-quantum-classical gravity",
^ manuscript in preparation) and “Methods” subsection “Discrete mas-
hDT1, π ϕ ðΦ, π Φ , xÞi =  hmðxÞi, ð39Þ
ter equation”), J(x, t) can involve higher order moments, and will gen-
erally be described by a jump process26,30. Its statistics can also depend
meaning that the back-reaction matrix Dμα 1, π Φ is nonvanishing. In the on the mass density, since in general the diffusion matrix Dμν 2, ij couples
“Methods” subsection “Newtonian limit of CQ theory” we give
to Lindblad operators. It is worth noting that the discrete CQ theories
examples of master equations for which Eq. (39) is satisfied, but their
considered in11,26,37 generically suppress higher order moments, and
details are irrelevant since we only require the expectation of the back-
often we expect that we can approximate the dynamics by a Gaussian
reaction force to be the expectation value of the mass—a necessary
process, but this need not be the case in general.
condition for the theory to reproduce Newtonian gravity.
The stochastic contribution to the Newtonian potential leads to
As a consequence of the coupling constant and observational
observational consequences which can be used to experimentally test
trade-offs derived in Eqs. (31) and (32), a non-zero D1, π Φ implies that
and constrain CQ theories of gravity for various choices of kernels
there must be diffusion in the momenta conjugate to πΦ. This diffusion
appearing in the CQ master equation. One immediate consequence is
is equivalent to adding a stochastic random process J(x, t) (the Lan-
that the variation in Newtonian potential leads to a variation of force
gevin picture), to the equation of motion (37) to give
experienced R 3 by a particle or composite mass via
!
F tot =  dR xmðxÞ∇ΦðxÞ. We can also estimate the time-averaged
∇2 Φ ΔT !
π_ Φ = ^ Jðt, xÞ,
 mðxÞ + uðΦ, mÞ ð40Þ force via ΔT 1
0 F tot where ΔT is the time resolution over which the
4πG force is measured and is the useful quantity when compared with
experiments. In the “Methods” subsection “Table-top experiments” we
^
where we allow some colouring to the noise via a function uðΦ, mÞ impose the constraint π ≈ 0 in Eq. (40) and find that the variance of the
^ (assumption
which can depend on Φ, and the matter distribution m magnitude of the time-averaged force experienced by a particle in a
(iv)). The noise process satisfies Newtonian potential is given by Eq. (146),

Em, Φ ½uJðx, tÞ = 0, Em, Φ ½uJðx, tÞuJðy, t 0 Þ = 2hD2 ðx, y, ΦÞiδðt, t 0 Þ, ð41Þ Z
2G2
σ 2F = d3 xd3 yd3 x 0 d3 y0 mðxÞmðyÞ
ΔT
where we have defined hD2 ðx, y, ΦÞi = Tr½Dμν y ð42Þ
2 ðx, y, ΦÞLμ ðxÞρLν ðyÞ, and ρ ! !0 ! !0
ðx  x Þð y  y Þ
is the quantum state for the decohered mass density. Here the m, Φ hD2 ðx 0 , y0 , ΦÞi,
jx  x 0 j3 jy  y0 j3
subscripts of Em, Φ allow for the possibility that the statistics of the
noise process can be dependent on the Newtonian potential and mass where the variation is averaged over the time resolution ΔT. We will use
distribution of the particle. The restriction on Em, Φ ½uJðx, tÞ follows this to estimate the variation in precision measurements of mass, such
from assumption (ii). If uJ(x, t) is Gaussian, Eq. (41) completely deter- as modern versions of the Cavendish experiment for various choices
mines the noise process, but in general, higher-order correlations are of hD2 ðx 0 , y0 , ΦÞi.
possible, although they need not concern us here, since we are only On the other hand, experimentally measured decoherence rates
interested in bounding the effects due to D2(x, y, Φ). can be related to D0. The important point is that the decoherence rate
In the non-relativistic limit, where c → ∞, we impose the momen- is dominated by the background Newtonian potential Φb due to the
tum constraint πΦ ≈ 0 and we recover Poisson’s equation for gravity, Earth. In the “Methods” subsection “Decoherence rates”, we show that
but with a stochastic contribution to the mass. This is precisely as for a mass whose quantum state is a superposition of two states jLi and
expected on purely physical grounds: in order to maintain coherence jRi of approximately orthogonal mass densities mL(x), mR(x), and
of any mass in superposition, there must be noise in the Newtonian whose separation we take to be larger than the correlation range of
potential and this must be such that we cannot tell which element of D0(x, y), the decoherence rate is given by
the superposition the particle will be in, meaning the Newtonian Z
potential should look like it is being sourced in part by a random mass 1
λ= dxdyDαβ y y
0 ðx, yÞðhLjLβ ðyÞLα ðxÞjLi + hRjLβ ðyÞLα ðxÞjRiÞ:
ð43Þ
distribution. In other words, the trade-off requires that the stochastic 2
component of the coupling obscures the amount of mass m at the
different points in space where the mass may be found. Via the coupling constant trade-off, Eqs. (42) and (43) then give
In the case where u is independent of Φ, it is simple to solve rise to a double-sided squeeze on the coupling D2. Equation (42) upper
Eq. (40) in terms of Green’s function for Poisson’s equation as in the bounds D2 in terms of the uncertainty of acceleration measurements
“Methods” subsection “Detecting gravitational diffusion”. A formal seen in gravitational torsion experiments, whilst the coupling constant
treatment of solutions to non-linear stochastic integrals of the more trade-off Eq. (43) lower bounds D2 in terms of experimentally mea-
general form of Eq. (40) can be found in ref. 82. A higher precision sured decoherence rates arising from interferometry experiments.
calculation would involve a full simulation of CQ dynamics, for We now show this for various choices of diffusion kernel, with the
example using unravelling methods26,78 or the path integral as in ref. 81. details given in the “Methods” subsection “Table-top experiments”.
However, care should be taken, as we have found that relativistic The bounds are summarised in Table 1. The diffusion coupling strength
corrections put constraints on the degree of diffusion even at low will be characterised by the coupling constant D2, which we take to be a
energy (J. Oppenheim and A. Russo, manuscript in preparation), and dimension-full quantity with units kg2 sm−3, and is related to the rate of
one should bear this in mind when drawing conclusions on the models diffusion for the conjugate momenta of the Newtonian potential. We
presented here. upper bound D2 by considering the variation of the time averaged

Nature Communications | (2023)14:7910 8


Article https://doi.org/10.1038/s41467-023-43348-2

Table 1 | Current experimental bounds on classical-quantum theories for different master equations and functional depen-
dence on the diffusion coefficient
Master equation Diffusion kernel Experimental squeeze
Continuous (ultra-local, non-rel.) D2(Φ; x, y) = D2(Φ)δ(x, y)D2(Φ) = ∑ncnΦn 10−41  D2  10−9 kg2 sm−3 (Eq. (44))
P
109  lP D2  1035 kg sm1 (Eq. (47))
2 2 2
Continuous (Eq. (116) or (118)) D2 ðΦ; x,yÞ =  lp D2 ðΦÞ∇2 δðx,yÞD2 ðΦÞ = n cn Φn
l2P P n n 1 l2P D2 25
Discrete (ultra-local) D2 ðΦ; x,yÞ = m D2 ðΦÞδðx,yÞD2 ðΦÞ = n c Φ 10  mP  10 kg (Eq. (46))
P

The diffusion coefficient is bounded from above by observed acceleration variations σ 2a seen in precision mass experiments via Eq. (42). In all cases the master equation is assumed to saturate the
bound which is used to find the lower bound the amount of diffusion on the quantum system by bounding D0 from coherence rates via Eq. (43). It is seen that minimally coupled continuous models
which are non-relativistic and do not create spatial correlations (we call these ultra-local) and have polynomial dependence on the Newtonian potential are ruled out, while continuous models with
non-local correlations, such as the Diosi–Penrose (DP) kernel of Eq. (116)or the relativity inspired kernel of Eq. (118), and ultra-local discrete models are less constrained. Here lP, mP are the Planck
length and Planck mass respectively, which are required in order for the dimensions of D2(Φ) to be the same in all cases.

acceleration σ a = σMF for a composite mass M which contains N atoms Taking a conservative bound on λ, for example, that arising from the
which we treat as spheres of constant density ρ with radius rN and mass interferometry experiment of59 which saw coherence in large organic
mN. We lower bound D2 via the coupling constant trade-off of Eq. (30) fullerene molecules with total mass Mλ = 10−24 kg over a timescale of
and then by considering bounds on the coherence time for composite 0.1s, gives an upper bound on the decoherence rate λ < 101 s−1. Full-
particles with total mass Mλ and which are made up of Nλ constituents, erene molecules are made up of Nλ ~ 103 particles with a typical atomic
each with typical length scale when in superposition Rλ and volume Vλ. size 10−15 m. After passing through the slits the molecule becomes
For continuous dynamics 〈D2(x, y, Φ)〉 = D2(x, y, Φ) since the dif- delocalised in the transverse direction on the order of 10−7 m before
fusion is not associated with any Lindblad operators. Let us now con- being detected. Since the interference effects are due to the super-
sider a very natural kernel, namely D2(x, y; Φ) = D2(Φ)δ(x, y) which is position in the transverse x direction, which is the direction of align-
both translation invariant, and does not create any correlations over ment of the gratings, it seems like a reasonable assumption to take the
space-like separated regions. We call dynamics which does not create size of the wavepacket in the remaining y, z direction to be the size of
correlations over space-like separated regions ultra-local since the- the fullerene, since we could imagine measuring the y, z directions
ories that are not of this form can still be non-signalling. This is a without effecting the coherence. We, therefore, take the volume
natural kernel from the point of view of constructing theories which Vλ ~ 10−1510−1510−7 m3 = 10−37 m3, which gives D2 ≥ 10−9 kg2 sm−3, and sug-
are diffeomorphism invariant. We also label this model as being non- gests that classical–quantum theories of Newtonian gravity with ultra-
relativistic since it does not include various relativistic corrections to local continuous noise are ruled out by experiment.
the diffusion. On the other hand, the discrete models appear less constrained
The decoherence rate for this kernel is found in the “Method” due to the suppression of the noise away from the mass density. For
subsection “Decoherence rates” and follows immediately from Eq. example consider the ultra-local discrete jumping models, such as the
(137). For a nucleon of mass Mλ and wavepacket volume Vλ, it is one given in the section “Discrete master equation” which have
qffiffiffiffi
λ = 2D0 M 2λ =V λ . In general, the squeeze will depend on the functional 3
l D ðΦ Þ
hD2 ðx, y, Φb Þi = P m2 b mðxÞ, where mP = _c G is the Planck mass and
choice of D2(Φ) on the Newtonian potential. However, in the presence qffiffiffiffi P

of a large background potential Φb, such as that of the Earth’s, we will l P = _Gc3
is the Planck length, required to ensure D2 has the units of
often be able to approximate D2(Φ) = D2(Φb). This is true for kernels kg2 sm−3. We find the squeeze on D2
that depend on Φ and ∇ Φ, though the approximation does not hold
for all kernels, for example D2 ~ −∇2Φ of Eq. (118) which creates diffu- 3
σ 2a Nr 4N ΔT lP M
sion only where there is mass density. For diffusion kernels D2(Φb) ≥ D ≥ λ, ð46Þ
mN G2 mP 2 λ
where the background potential is dominant, we find the promised
squeeze on D2(Φb)
and plugging in the numbers tells us that discrete theories of classical
σ 2a Nr 4N ΔT N M2 gravity are 3 not ruled out by experiments and we
≥ D2 ≥ λ λ , ð44Þ
find 101 kg ≥ mP D2 ≥ 1025 kg.
l
V b G2 V λλ P
We can also consider other noise kernels, with examples and some
where Vb is the volume of space over which the background Newtonian discussion is given in the section “Examples of Kernels saturating the
potential is significant. Vb enters since the variation in acceleration is decoherence diffusion coupling constants trade-off”. A natural kernel
2
found to be is D2 ðx, y, Φb Þ =  l P DðΦb Þ∇2 δðx, yÞ. The inverse Lindbladian kernel
satisfying the coupling constants trade-off is to zeroeth order in Φ(x),
Z 0 ðΦb Þ
D2 G2 the Diosi-Penrose kernel D0 ðx, y, Φb Þ = Djxyj . We here consider higher-
d x 0 D2 ðΦb Þ,
3
σ 2a ∼ ð45Þ
r 4N NΔT order terms such as those coming from the relativistic theory and in
particular the diffusion kernel of Eq. (118). For this choice of dynamics,
3
where the d x 0 integral is over all space. we find the squeeze for D2 in terms of the variation in acceleration
This immediately rules out continuous theories of Newtonian
gravity with noise everywhere, i.e., with a diffusion coefficient inde- σ 2a Nr 3N ΔT 2 N λ M 2λ
≥ l P D2 ≥ : ð47Þ
pendent of the Newtonian potential, since the integral will diverge. We G 2 Rλ λ
consider the relativistic case elsewhere.
Standard Cavendish-type classical torsion balance experiments49 Using the same numbers as for the ultra-local continuous model,
with Rλ ∼ V λ ∼ 1012 m we find that classical torsion experiments
1=3
measure accelerations of the order 10−7 m s−2 over minutes ΔT ~ 102, so
upper bound D2 by 109 kg sm1 ≥ l P D2 , whilst interferometry
2 2
a very conservative bound is σa ~ 10−7 m s−2, whilst for a kg mass N ~ 1026
experiments bound D2 from below via l P D2 ≥ 1035 kg sm1 .
2 2
and rN ~ 10−15 m. Conservatively taking V b ∼ r 2E h m3 where rE is the
radius of the Earth and h is the atmospheric height gives Equations (44), (46) and (47) show that classical theories of gravity
D2 ≤ 10−41 kg2 sm−3. The decoherence rate λ is bounded by various are squeezed by experiments from both ways. We have here been
experiments83. Typically, the goal of such experiments is to witness extremely conservative, and we anticipate that further analysis, as well
interference patterns of molecules that are as massive as possible. as near-term experiments, can tighten the bounds by orders of

Nature Communications | (2023)14:7910 9


Article https://doi.org/10.1038/s41467-023-43348-2

magnitude. There are several proposals for table-top experiments to positive throughout the dynamics. We have shown that any theory
precisely measure gravity, some of which have recently been per- which preserves probabilities and treats one system classically is
formed, and which could give rise to tighter upper bounds on D2. Some required to have fundamental decoherence of the quantum system,
of these experiments involve millimeter-sized masses whose gravita- and diffusion in phase space, both of which are signatures of infor-
tional coupling is measured via torsional pendula61,62, or rotating mation loss. Using a CQ version of the Kramers–Moyal expansion, we
attractors63. With such devices, the gravitational coupling between have derived a trade-off between decoherence on the quantum sys-
small masses can be measured while limiting the amount of other tem, and the system’s diffusion in phase space. The trade-off is
sources of noise. There are proposals for further mitigating the noise expressed in terms of the strength of the back-reaction of the quantum
due to the environment, including inertial noise, gas particle collisions, system on the classical one. We have derived the trade-off both in
photon scattering on the masses, and curvature fluctuations due to terms of coupling constants of the theory and in terms of observa-
other sources84–86. Other experiments are based on interference tional quantities that can be measured experimentally.
between masses; for example, atomic interferometers allow for the In the case of gravity, the trade-off places a lower bound on the
measurement of the curvature of space-time over a macroscopic rate of diffusion of the gravitational degrees of freedom in terms of the
superposition87–89. decoherence rate of particles in superposition. We find that theories
We can get stronger lower bounds via improved coherence that treat gravity as fundamentally classical, are not ruled out by cur-
experiments. Typically, the goal of such experiments is to witness rent experiments, however, we have been able to rule out a broad
interference patterns of molecules that are as massive as possible, parameter space of Newtonian theories. This is done partly through
while here, we see that the experimental bound on CQ theories is table-top observations via Eqs. (44), (46) and (47). Given any diffusion
generically obtained by maximising the coherence time for massive kernel, we can compute the inaccuracy of mass measurements due to
particles with as small wave-packet size Vλ. fluctuations in the gravitational field, and using the trade-off, we can
Thus far we have considered local effects on particles due to derive a bound on the associated decoherence rate. This allows us to
diffusion. While this enables us to rule out some types of theories, the rule out broad classes of theories in terms of their diffusion kernel. For
bounds are generally weak if one wants to rule out all of them. How- example, we are able to rule out a number of non-relativistic theories
ever, it may be possible to do so via cosmological considerations. In which back-react continuously in phase space.
attempting to place experimental constraints on this diffusion, it is Any theory that treats gravity classically has fairly limited freedom
also worth considering other regimes, such as longer range effects to evade the effects of the trade-off. There is the freedom to choose the
which might be detected by gravitational wave detectors such as LIGO, diffusion or decoherence kernels D2 ðx, x 0 Þ and D0 ðx, x 0 Þ, but the trade-
or table-top interferometers90,91. We leave a detailed study of the effect off restricts one in terms of the other. Then, because of the results
of gravitational diffusion on cosmological scales and LIGO to future proven in30, one can consider two classes of theory, those which are
work. It suffices to mention that the effect will again depend on the continuous realisations and whose diffusion can only depend on the
form of the kernel D2 ðx, x 0 Þ. Our estimates (J. Oppenheim and Z. Weller- gravitational degrees of freedom, and discrete theories whose diffu-
Davies, “Estimating space-time diffusion in interferometers", unpub- sion can also depend directly on the matter fields. Examples of both
lished note) suggest that local effects from table-top experiments classes of the theory are given in the “Methods” subsection “Newtonian
currently place a stronger bound on gravitational theories than LIGO limit of CQ theory”. Finally, one could consider theories that do not
currently does. In particular, unlike gravitational wave measurements, reproduce the weak field limit of General Relativity to all distances,
which are reasonably high-frequency events requiring extraordinarily namely we could imagine that the interaction Hamiltonian of Eq. (36)
high precision in the relative displacement of the arm length from its does not hold to arbitrarily short distances, or arbitrarily high mass
average, it is preferential to have a lower precision measurement, densities. This is reasonable since we do expect the Newtonian theory
which occurs over a longer time period to allows for the diffusion in to break down at short distances where relativistic corrections at high
path length to build up, and with a smaller uncertainty in the average energy affect the low-energy behaviour of the theory. One could also
length of the arm itself. Furthermore, since the LIGO arm is kept in a consider modifying D1 ðx, x 0 Þ in some other way, for example, by
vacuum, we do not expect strong bounds on discrete models where making it slightly non-local, or by disallowing arbitrarily high mass
the diffusion is associated with an energy density. densities, or by including an additional contribution such as the fric-
tion term discussed in the continuous master equation. All of these
Discussion modifications would seem to violate Lorentz invariance in some way,
A number of direct proposals to test the quantum nature of gravity are and likely lead to observational consequences93.
expected to come online in the next decade or two. These are based on Here, we have only given an order of magnitude estimate of when
the detection of entanglement between mesoscopic masses inside gravitational diffusion will lead to appreciable deviations from New-
matter-wave interferometers64–70,72,73. For these experiments, some tonian gravity. The most promising experiments bounding the diffu-
theoretical assumptions are needed: one requires that it is only grav- sion appear to be table-top experiments which precisely measure the
itons that travel between the two masses and mediate the creation of mass of an object. This is an area that is important from the perspective
entanglement. If this is the case, then the onset of entanglement of weight standards, for example, those undertaken by NIST on the 1 kg
implies that gravity is not a classical field. These can be thought of as mass standard K20 and K494. The increased precision and measuring
experiments that if successful, would confirm the quantum nature of time of Kibble Balances95 and atomic interferometers87,88,96,97 would
gravity (although other alternatives to quantum theory are possible92). make such measurements an ideal testing ground, both to further
Here, we come from the other direction, by supposing that gravity constrain the diffusion kernel and to look for diffusion effects, whose
is instead classical, and then exploring the consequences. Theories in dependence on the test mass is outlined in the “Methods” subsection
which gravity is fundamentally classical were thought to have been “Detecting gravitational diffusion”. Here, we have found that the
ruled out by various no-go theorems and conceptual difficulties. resolution time ΔT over which variations of acceleration are estimated
However, these no-go theorems are avoided if one allows for non- affects the strength of the bound, and it would be helpful if future
deterministic coupling as in11,21–26,30,37. We have here proven that this experiments reported this value. Since we have found that CQ theories
feature is indeed necessary and made it quantitative by exploring the predict an uncertainty in mass measurements it is perhaps intriguing
consequences of complete positivity on any dynamics that couples that different experiments to measure Newton’s constant G yield
quantum and classical degrees of freedom. Complete positivity is results whose relative uncertainty differs by as much as
required to ensure the probabilities of measurement outcomes remain 5 × 10−4 m3 kg−1 s−2, which is more than an order of magnitude larger

Nature Communications | (2023)14:7910 10


Article https://doi.org/10.1038/s41467-023-43348-2

than the average reported uncertainty52–54. If one were to try and by Oppenheim11. In other words, one might expect some gravitational
explain the discrepancy in G measurements via gravitational diffusion, diffusion, because, from an effective theory point of view, one is in a
then for all the kernels we studied in Section ’Physical constraints on regime where space-time is behaving classically. There are even claims
the classicality of gravity’ we find that the variation in acceleration that holographic effects could cause stochasticity113–115 in the gravita-
depends on p1ffiffiffi N
the number of nucleons in the test mass, so that masses tional field. However, the trade-off we have derived is a direct con-
with smaller volume should yield larger uncertainty and this would be sequence of treating the background space-time as fundamentally
the effect to look for in measurement discrepancies. The relatively classical. In a fully quantum theory of gravity, the interaction of the
large uncertainty in such measurements, also makes it challenging for gravitational field with particles in a superposition of two trajectories
table-top experiments to place strong upper bounds on gravitational will cause decoherence, but coherence can then be restored when the
diffusion. two trajectories converge. This is because the particle’s position is
Turning to the other side of the trade-off, improved decoherence entangled with the gravitational field (or dressed by it), and this
times would further squeeze theories in which gravity remains classi- entanglement is erased when the different paths of the superposition
cal. While a current experimental challenge is to demonstrate inter- converge. This is what happens when electrons interact with the elec-
ference patterns using larger and larger mass particles, we here find tromagnetic field while passing through a diffraction grating, yet still
the bounds in Eqs. (44) and (46) depend on the expectation of the form an interference pattern at the screen. This is a non-Markovian
particle’s mass density in ways that depend on the particular kernel. effect—the which-path superposition decoheres almost immediately,
Thus interference experiments with particles of high mass density but this is false-decoherence116 so the amount of diffusion can be arbi-
rather than mass can be preferable. There are also kernels, for which trarily small and is unrelated to the coherence time of the superposition.
the relevant quantity is the expectation of the mass density or M 2λ =V λ , On the other hand, the trade-off we derived is a direct con-
which will depend on both the particle’s mass Mλ and volume Vλ of the sequence of the positivity condition, which is a direct consequence of
wave-packet used in the interference experiment, a quantity which is the Markovian assumption. In the non-Markovian theory where Gen-
not always obtainable from many reports on such experiments. While eral Relativity is treated classically, one still expects the master equa-
this dependence might initially appear counter-intuitive, it follows tion to take the form found in11, but without the matrix whose elements
from the fact that in order to relate the trade-off in terms of coupling are Dμνn needing to be positive semi-definite at all times
117,118
. It would
constants to observational quantities, and in particular, the deco- therefore be surprising, if a quantum theory of gravity predicted
herence rate, we took expectation values of the relevant quantities to anything close to the level of diffusion predicted by the decoherence-
get a trade-off in terms of only averages. And indeed the decoherence vs-diffusion trade-off, as there would be no need for diffusion to
rate, which is an expectation value, can easily depend on the wave- explain the coherence of superpositions. The regime in which the
packet density, as we see from examples in the section “Deco- classical–quantum theory can be regarded as an effective one is taken
herence rates”. up in ref. 119, both to address the issue of false decoherence, and also
Since we here show that all theories that treat gravity classically to explore the regime in which the classical–quantum theory may be a
necessarily decohere the quantum system, another constraint on useful tool to understand the back-reaction of quantum matter in
theories that treat gravity classically is given by constraints on funda- space–time, such as during black-hole evaporation, and during infla-
mental decoherence. These are usually constrained by bounds on tion. If we instead regard the theory as describing a fundamentally
anomalous heating of the quantum system98. However, these con- classical space-time, then it follows from the decoherence-diffusion
straints are not in themselves very strong, since fundamental deco- trade-off, that the diffusion is either fundamental or its source is not
herence effects can be made arbitrarily weak. In the simplified model in describable within quantum or classical mechanics (J. Oppenheim,
the “Methods” subsection “Newtonian limit of CQ theory”, the strength “Post-quantum soup", unpublished note).
of the decoherence depends on the strength of the gravitational field,
thus, constraints due to heating98–111 can be suppressed, either by Methods
scaling the Lindbladian coupling constants or by having strong deco- Positivity conditions and the trade-off between decoherence
herence effects more pronounced near stronger gravitational fields and diffusion
such as near black holes where one expects information loss to occur. In this section, we will introduce two forms of positivity conditions
The necessity for decoherence to heat the quantum system is further used to prove the decoherence diffusion trade-off.
weakened by the fact that the dynamics are not Markovian on the The first inequality we would like to introduce is
quantum fields, if one integrates out the classical degrees of freedom, Z
space-time acts as a memory. This potentially captures some of the
dzA*μ ðz, z 0 ÞΛμν ðzjz 0 , δtÞAν ðz, z 0 Þ ≥ 0, ð48Þ
non-Markovian features advocated in ref. 112, who recognised that
Markovianity is a key assumption in attempts to rule out fundamental
decoherence or information loss. Here, however, we see that there is which holds for any Aμ ðz, z 0 Þ for which Eq. (48) is well defined: i.e., so
less freedom than one might imagine. If the Lindbladian coupling that the distributional derivatives in Eq. (48) are well defined.
constants are made small to reduce direct heating, the gravitational We can derive the positivity condition (48) from the positivity of
diffusion must be large. Thus, heating constraints which place bounds Λμν ðzjz 0 Þ, which must be a positive semi-definite matrix in μν. More
μ
on D0 ðx, x 0 Þ place additional constraints on D2 ðx, x 0 Þ. In35, it was found precisely, the eigenvalues of Λμν ðzjz 0 Þ, which we denote by λ ðzjz 0 Þ
that for the Newtonian models of ref. 33, large D2 ðx, x 0 Þ creates sec- must be positive. They must be positive in the distributional sense,
μ
ondary heating which further constrain the theory experimentally. The since we allow for the case that λ ðzjz 0 Þ is a positive distribution, for
0
decoherence-diffusion trade-off implies that this is a general feature of example λ ðzjz 0 Þ ∼ δðz  z 0 Þ. Hence we require
all theories which treat gravity classically. Z
While the absence of diffusion could rule out theories where μ
dzdz 0 λ ðzjz 0 ÞPðz, z 0 Þ ð49Þ
gravity is fundamentally classical, the presence of such deviations, at
least on short time scales, might not by itself be a confirmation of the
classical nature of gravity. Such effects could instead be caused by is positive for any positive smearing function Pðz, z 0 Þ. Since each λμ
quantum theories of gravity whose classical limit is effectively described must be positive, we can also pick a different smearing function for

Nature Communications | (2023)14:7910 11


Article https://doi.org/10.1038/s41467-023-43348-2

each μ, so that Since we know D2 and D0 must be positive semi-definite, we know from
Z Schur decomposition that
μ
dzdz 0 λ ðzjz 0 ÞP μ ðz, z 0 Þ ð50Þ
1 bry
2D2 kDbr
1 D0 D1 ,
ð57Þ

should be positive for any vector P μ ðz, z 0 Þ with all positive entries. We and ðI  D0 D0  1ÞDbr1 = 0, where D0 is the generalised inverse of D0.
can then write the matrix Λμν ðzjz 0 Þ in terms of its eigenvalues Furthermore, if D0 vanishes, then clearly Dbr
1 must also vanish in order
for (56) to be positive semi-definite.
ρ
Λμν ðzjz 0 Þ = U μy 0 0 ν 0
ρ ðzjz Þλ ðzjz ÞU ρ ðzjz Þ: ð51Þ
Classical-quantum dynamics with fields
We can then see the positivity of Eq. (48) directly since In this section, we describe CQ dynamics in the case where the Lind-
Z Z blad operators and the phase-space degrees of freedom can have
μ spatial dependence z(x), Lμ(x).
dzA*μ ðz, z 0 ÞΛμν ðzjz 0 , δtÞAν ðz, z 0 Þ = dzðUAÞyμ ðzjz 0 Þλ ðzjz 0 ÞðUAÞμ ðz, z 0 Þ
Z For the case of fields, operators O(x) constructed out of local
μ fields ϕ(x) will in general be unbounded and hence the Stinespring
= dzjðUAÞμ j2 ðz, z 0 Þλ ðzjz 0 Þ
dilation theorem does not hold. This problem is a common one in the
ð52Þ study of algebraic quantum field theory and we can get around it by
considering the case in which operators of interest are obtained by
which is positive as a consequence of Eq. (50). smearing the local fields over bounded functionals F. For example, we
As a consequence of Eq. (48) being positive, we also know that can first smear the local field fields over a smearing function
Z  f, ϕf = ∫dxϕ(x)f(x) and then consider bounded functions of ϕf such as
Fðϕf Þ = eiϕf . In doing this we can write a CQ version of the Stinespring
Tr dzΛμν ðzjz 0 ÞOμ ðz, z 0 Þρðz 0 ÞOyν ðz, z 0 Þ ≥ 0 ð53Þ
dilation theorem exactly and proceed along the lines of Oppenheim11
to show that any completely positive CQ map can be written in the
will be positive for any vector of operators (potentially phase space form
dependent) Oμ ðz, z 0 Þ. This follows from the cyclicity of the trace and Z
the fact that Λμν ðzjz 0 ÞOyν ðz, z 0 ÞOμ ðz, z 0 Þ will be a positive operator so
ρ0 ðzÞ = dzdxdyΛμν ðzjz 0 ; x, yÞLμ ðx, z, z 0 Þϱðz 0 ÞLyν ðy, z, z 0 Þ, ð58Þ
long as Eq. (48) holds. A common choice of Oμ would be the Lindblad
operator Lμ appearing in the master equation.
The inequality in Eq. (48) proves useful to derive positivity where the positivity condition states
conditions on the coupling constants appearing in the master Z
equation, whilst Eq. (53) is useful in deriving the observational
dzdxdyA*μ ðx, z, z 0 ÞΛμν ðzjz 0 ; x, yÞAν ðy, z, z 0 Þ ≥ 0: ð59Þ
trade-off for the continuous master equation as we shall now
discuss.
We shall assume that we deal with dynamics which can be written
General trade-off between decoherence and diffusion in Lindblad form, as is usually assumed in the unbounded case120.
coefficients
We can get a general trade-off between the decoherence and diffusion CQ Kramers–Moyal expansion for fields
coefficients which appear in the master equation, arriving at a trade-off Just as in the section “The CQ Kramers–Moyal expansion”, we can
between the decoherence and diffusion coefficients in terms of the formally introduce the moments of the transition amplitude
back-reaction drift coefficient Dμα
1, i . Z
Consider Eq. (48), and choose Aμ = δ αμ aα + bμ ðz  z 0 Þi . By inte-
i
M μν
n, i ...i ðw1 , . . . wn ; x, y, δtÞ = DzΛμν ðzjz 0 ; x, y, δtÞ ðz  z 0 Þi1 ðw1 Þ . . . ðz  z 0 Þin ðwn Þ
grating parts over the phase space degrees of freedom, we find
1 n

ð60Þ
2bμ Dμν + bμ Dμβ + a*α Dαμ + a*α Dαβ
i* j i* i
2, ij bν 1, i aβ 1, i bμ 0 aβ ≥ 0: ð54Þ
which we assume to exist; which might involve a smearing of the
Taking i ∈ {1, …, n}α ∈ {1, …, p} and μ ∈ {1, …, p + 1}, we can write operators z(x). Defining L0 ðxÞ = δðxÞI, we can define the coefficients
this as a matrix positivity condition Dμν
n, i ...i implicitly via
1 n

" #  μ ν
* 2D2 Dbr
1 b M μν 0 μν
n, i ...i ðz , w1 , . . . wn ; x, y, δtÞ = δ 0 δ 0 + δtn!Dn, i ...i ðw1 , . . . wn ; x, y, δtÞ:
½b , a*  ≥0 ð55Þ 1 n 1 n

Dbr
1 D0 a ð61Þ

where D2 is the (p + 1)n × (p + 1)n matrix with elements Dμν br


2, ij , D1 is The characteristic function then takes the form
the (p + 1)n × p matrix with rows labelled by μi and columns Z R
labelled by β with elements Dμβ 1, i and D0 is the p × p decoherence dwuðwÞðzðwÞz 0 ðwÞÞ μν
C μν ðu, z 0 ; x, yÞ = Dzei Λ ðzjz 0 ; x, yÞ ð62Þ
matrix with elements Dαβ br
0 . D1, i describes the quantum back-
reacting components of the drift. Equation (55) is equivalent to
the condition that the ððp + 1Þn + pÞ × ððp + 1Þn + pÞ matrix and expanding out the exponential this takes the form
" # 1 Z
X
2D2 Dbr ui ðw1 Þ . . . uin ðwn Þ
1
k0: ð56Þ C μν ðu, z 0 ; x, yÞ = dw1 . . . dwn M μν 0
n, i ...i ðz , w1 , . . . wn ; x, y, δtÞ
n! 1 n

Dbr
1 D0 n=0

ð63Þ

performing the inverse Fourier transform, allows us to write the


transition amplitude in terms of functional derivatives of the delta

Nature Communications | (2023)14:7910 12


Article https://doi.org/10.1038/s41467-023-43348-2

function Since we know D2 and D0 must be positive semi-definite, we know


from Schur decomposition that
1 Z
X M μν 0
n, i ...i ðz , w1 , . . . wn ; x, y, δtÞ
n
δ
Λμν ðzjz 0 ; x, y, δtÞ = dw1 . . . dwn 1 n
δðz, z 0 Þ
n! δz 0i ðw1 Þ . . . z 0i ðwn Þ 1 bry
ð69Þ
n=0 1 n 2D2 kDbr
1 D0 D1
ð64Þ
and
and we can use this to write a CQ master equation in the form

1 Z  
ðI  D0 D1 br
0 ÞD1 = 0,
ð70Þ
∂ϱðz, δtÞ X
n
δ
= dw1 . . . dwn ð1Þn D00 ðz, w1 , . . . wn ÞϱðzÞ  i½H, ϱðzÞ
∂t δz i1 ðw1 Þ . . . z in ðwn Þ n, i1 ...in
n=1
Z
1 αβ
where D1
0 is the generalised inverse of D0. Furthermore, from Equa-
+ dxdyDαβ y
0 ðz; x, yÞLα ðxÞϱðzÞLβ ðyÞ  D0 ðz; x, yÞfLβ ðyÞLα ðxÞ, ϱg tion (70), we see if D0 vanishes, then clearly Dbr
2 1 must also vanish in
X1 X Z
+ dxdydw1 . . . dwn ð1Þn
order for (68) to be positive semi-definite.
n = 0 μν≠00
 
δ n
Dμν ðz, w1 , . . . wn ; x, yÞLμ ðxÞϱðzÞLyν ðyÞ :
Observational trade-off in the presence of fields
δz i1 ðw1 Þ . . . z in ðwn Þ n, i1 ...in
We can use the same methods to arrive at an observational trade-off
ð65Þ using the field-theoretic version of the Cauchy–Schwartz inequality in
Eq. (18). This arises from the positivity of
Since we are interested in studying dynamics with local back-
reaction, we shall hereby take Dμν μν Z 
1 ðz, w; x, yÞ = D1 ðxÞδðx, yÞδðx, wÞ.
By the decoherence diffusion trade-off, which we derive in the Tr dzdz 0 dxdyΛμν ðzjz 0 , x, yÞOμ ðz, z 0 , xÞρðz 0 ÞOyν ðz, z 0 , yÞ ≥ 0 ð71Þ
next subsection, this also means that the diffusion matrix
Dμν
2, ij ðz, w1 , w2 , x, yÞ is lower bounded by the matrix for any local vector of CQ operators Oμ ðz, z 0 , xÞ. We have to be careful,
Dμα 1 βν*
1 ðxÞðD0 Þαβ ðx, yÞD1 ðyÞδðw1 , xÞδðw2 , yÞ. This can be R seen more pre- since (71) is not in general well defined since Oμ may not be trace-class.
α i
cisely, by taking Eq. (59) with Aμ ðxÞ = δ μ aα ðxÞ + dwbμ ðx, wÞðz  We hence assume that we consider states ρ(z) and operators Oμ ðz, z 0 , xÞ
0
z Þðx, wÞ and applying the same methods as in the subsection “Trade- for which Eq. (71) is well defined. Since we are interested in getting an
off between diffusion and decoherence couplings in the presence of observational trade-off we expect this to always be the case for
fields”. Without loss of generality we thus take physical classical–quantum states ρ(z).
D2(z, w1, w2, x, y) = D2(z, x, y)δ(x, w1)δ(y, w2). We shall use Eq. (71) to arrive at a (pseudo) inner product on a
vector of operators Oμ via
Trade-off between diffusion and decoherence couplings in the
presence of fields Z h i
 ,O
hO  i= dzdz 0 dxdyTr Λμν ðzjz 0 x, yÞO1μ ðxÞϱðz 0 ÞOy2ν ðyÞ ð72Þ
In the field-theoretic case, the positivity condition is given by Eq. (59) 1 2
and we can find a trade-off between decoherence and diffusion by
α R i pffiffiffiffiffiffiffiffiffiffiffiffiffi
considering Aμ ðxÞ = δ μ aα ðxÞ + dxbμ ðxÞðz  z 0 ÞðxÞ. So that  = hO,  Oi  ≥ 0 due to Eq. (71). Technically this is not positive
where jjOjj
Z definite, but again, this will not worry us. Hence, so long as jjO  jj≠0,
2
dxdy2bμ ðxÞDμν μβ αμ
i* j i* * i
2, ij ðx, yÞbν ðyÞ + bμ ðxÞD1, i ðx, yÞaβ ðyÞ + aα ðxÞD1, i ðx, yÞbμ ðyÞ which holds due to the CQ inequality derived in the derivation of the
Pawula theorem30, we again have a Cauchy–Schwartz inequality
+ a*α ðxÞDαβ
0 ðx, yÞaβ ðyÞ ≥ 0
ð66Þ jjO  jj2  jhO
 jj2 jjO  ,O
 ij2 ≥ 0: ð73Þ
1 2 1 2

where we use the shorthand notation :¼ and Dμν


2, ij ðz, x, yÞ Dμν
2, ij ðx, yÞ α
Choosing O1, μ ðxÞ = δ μ Lα ðxÞ and
similarly Dαμ ðz; x, yÞ :¼ Dαμ
ðx, yÞ. R i
1, i 1, i
O2, μ ðxÞ = dx 0 b ðxÞðz  z 0 Þi ðxÞLμ ðxÞ, one finds
Taking i ∈ {1, …, n}α ∈ {1, …, p} and μ ∈ {1, …, p + 1}, we can write
this as a matrix positivity condition Z h i
 jj2 =
jjO dzdxdyTr Dαβ y
" # 1 0 ðz; x, yÞLα ðxÞϱðzÞLβ ðyÞ :¼ hD0 i ð74aÞ

R * 2D2 ðx, yÞ Dbr
1 ðx, yÞ bðyÞ
dxdy½b ðxÞ, a*ðxÞ ≥0 ð67Þ
Dbr
1 ðx, yÞ D0 ðx, yÞ αðyÞ Z h i
 jj2 = 2
jjO dzdxdyTr b ðxÞDμν
j* y i
where D2(x, y) is the (p + 1)n × (p + 1)n matrix-kernel with elements 2 2, ij ðz; x, yÞLμ ðxÞϱðzÞLν ðyÞb ðyÞ ð74bÞ
Dμν br
2, ij ðx, yÞ, D1 ðx, yÞ is the (p + 1)n × p matrix-kernel with rows labelled by
μi and columns labelled by β with elements Dμβ 1, i ðx, yÞ and D0(x, y) is the Z h i2  Z 2
 
p × p decoherence matrix-kernel with elements Dαβ br
0 ðx, yÞ. D1, i describes
 ,O
jhO 1
 ij2 =  dzdxTr bi* ðxÞDαν ðz; xÞL ðxÞϱðzÞLy ðxÞ  :¼ 
2  1, i α ν  
i*
dxb ðxÞDbr
1, i ðxÞ


the quantum back-reacting components of the drift.
ð74cÞ
Equation (67) is equivalent to the condition that the
ððp + 1Þn + pÞ × ððp + 1Þn + pÞ matrix of operators
Taking the limit b ðxÞ ! δðx, x Þb ðx Þ, we arrive at a local trade-
i i
" #
2D2 Dbr off between diffusion, drift and total decoherence. In particular,
1
k0 ð68Þ using (74), the definitions of the expectation values of couplings
Dbr
1 D0
defined in Eq. (33) and the fact that for back-reaction the
be positive semi-definite. Here we are viewing the objects of expectation value of D0 cannot vanish, we arrive at the observa-
Eq. (68) as matrix-kernels, so that for any position-dependent tional trade-off of Eq. (34)
μ R
vector bμ ðxÞ, ðD2 bÞi ðxÞ = dyDμν
i j
2, ij ðx, yÞbν ðyÞ. h i
b ðx Þ 2hD2, ij ðx , x ÞihD0 i  jhDbr  2 j 
i
1, i ðx Þij b ðx Þ ≥ 0 ð75Þ

Nature Communications | (2023)14:7910 13


Article https://doi.org/10.1038/s41467-023-43348-2

which we write in matrix form as coupled matter


Z  
2hD2 ðx , x ÞihD0 ikhDbr  br  y
ð76Þ ∂g ij ∂ϕ
1 ðx ÞihD1 ðx Þi : S= d4 x π ij + π m m  NH  N i Hi , ð84Þ
∂t ∂t
It then follows directly from Eq. (76) that when the back-reaction
αβ
is sourced by either D0μ 1, i or D1, i components we can arrive at the where we are ignoring the boundary contributions to the action. Here,
observational trade-off in terms of the total drift
  
c4 16πG 1 1
y H  g 1=2 R + g ik g jl π ij π kl  π 2 + HðmÞ , ð85Þ
8hD2 ðx , x ÞihD0 ikhDT1 ðx ÞihDT1 ðx Þi , ð77Þ 16πG 2
c g 1=2 2
y
where in Eq. (77) recall that the definition of hDT1 ðx Þi is given by Eq. (35)
c3
in the main body. Hi  g ∇ π jk + HðmÞ
i ,
ð86Þ
8πG ij k
A spatially averaged observational trade-off are the Hamiltonian and momentum constraints and πij is defined in
It is also useful to note that one can arrive at a spatially averaged terms of the extrinsic curvature tensor of constant t surfaces, Kij, via
observational trade-off which can be used to bound all of the elements
of the diffusion matrix, not just its diagonals. Specifically, taking Eq. c3  
π ij   g 1=2 K ij  Kg ij : ð87Þ
(74) with bi(x) = bi a constant, we arrive at the trade-off 16πG
Z Z Z y
It is useful to note that the matter densities HðmÞ , HiðmÞ can be
8 dxdyhD2 ðx, yÞihD0 ik dxDT1 ðxÞ dxDT1 ðxÞ , ð78Þ
related to the matter stress-energy Tμν via
pffiffiffi 2 00
where we define the expectation matrix HðmÞ = gN T , ð88aÞ
Z h i
hD2 ðx, yÞiij = dzTr Dμν y pffiffiffi 0
2, ij ðz; x, yÞLμ ðxÞϱðzÞLν ðyÞ : ð79Þ HiðmÞ = g NT i : ð88bÞ

For the Newtonian limit discussed in the main body this bounds We here take the Newtonian limit of the gravitational field to be
the diffusion in terms of the total mass of the particle given by

Z    
M2 Φ 2Φ c2
dxdyhD2 ðx, yÞi ≥ : ð80Þ N = 1 + 2 , N i = 0, g ij = 1  2 δ ij , π ij =  π Φ δ ij , ð89Þ
16λ c c 6

We can also arrive at a trade-off in terms of the effective New- with Φ(x) corresponding to the Newtonian potential. The choice of
i 2 ij
tonian potential sourced by the masses by taking b ðxÞ = jxxj
1
. In this π ij =  c6 π Φ δ , is to ensure that πΦ is canonically conjugate to Φ. A
case, we find the trade-off more detailed derivation starts from the full weak field metric, and
gauge fixing of the shift can be found in81. Here, we find the effective
Z *Z +*Z +y
hD2 ðx, yÞi DT1 ðxÞ DT1 ðxÞ action can be written
8 dxdy hD ik dx dx
jx  xjjx  yj 0 jx  xj jx  xj Z  
4 ∂Φ ∂ϕ
ð81Þ S= d x πΦ + π m m  H Newt , ð90Þ
∂t ∂t

which for the Newtonian limit gives a trade-off between the diffusion where the Newtonian Hamiltonian is given by
matrix and the effective Newtonian potential of the particle as sourced
by its expectation value H Newt = H c + H 0ðmÞ + H I ð91Þ

Z R 2
hD2, πΦ π Φ ðx, yÞi  dx hjmðxÞi 
^
x xj
^ x Þij2
jhΦð with
dxdy ≥ = , ð82Þ
jx  xjjx  yj 16λ 16G2 λ Z !
3 2Gπc2 2 ð∇ΦÞ2
where we have defined the effective Newtonian potential Hc = d x  πΦ + ð92Þ
R 3 8πG
^ =  G dx hmðxÞi
as hΦi
^
jx xj .
the pure gravity Hamiltonian, and
Newtonian limit of CQ theory Z
In this section we motivate the Newtonian limit of gravity used in
HI = d3 xΦðxÞmðxÞ ð93Þ
Section ’Physical constraints on the classicality of gravity’121. A fuller
treatment can be found in81. We begin with classical General Relativity
in the ADM formulation36. To derive the Hamiltonian, we start from the is the interaction Hamiltonian, from which we see that non-relativistic
3+1 split of the four metric matter couples to the Newtonian potential through its mass density
   m(x). In the case where we have the state of matter being described by a
2
ds2 =  ðNc dtÞ + g ij dx i + N i c dt dx j + N j c dt , ð83Þ point particle δ(x−x(t)) of mass m the pure matter Hamiltonian would be

in which case, denoting ϕm, πm as canonical variables for the ij


δ pi pj
H 0ðmÞ = mc2 + : ð94Þ
matter degrees of freedom, we can write the action for minimally 2m

Nature Communications | (2023)14:7910 14


Article https://doi.org/10.1038/s41467-023-43348-2

The equations of motion for the gravitational degrees of freedom generally not equivalent to taking the master equations of
reads Oppenheim11, and then taking the appropriate limit81.

_ =  4πGc π ,
2 Continuous master equation
Φ Φ
ð95Þ
3 For the class of master equations with continuous back-reaction,
specifying that the first moment on average satisfies Eq. (100) is
enough (up to drift terms which vanish under trace) to ensure the
∇2 Φ ð96Þ
π_ Φ =  mðxÞ, master Equation includes a term
4πG
Z  
∂ϱ δϱ δϱ
which, for πΦ = 0 yields the Newtonian solution for a stationary mass ≈ fH c ðΦÞ, ϱg  i½H 0ðmÞ , ϱ + ^
d3 x mðxÞ + ^
mðxÞ
density. In a Louivile formulation the dynamics for the density ∂t δπ Φ δπ Φ
Z 2
ρ(Φ, πΦ, xi, pi) is given by δ ð101Þ
+ d3 xd3 y ðD ðΦ, π Φ ; x, yÞϱÞ
δπ Φ ðxÞδπ Φ ðx 0 Þ 2
Z Z
∂ρ ∂ρ δρ 1  
= fH c + H 0ðmÞ , ρg  ∂i ΦðxÞ
3
+ d x mðxÞ , ð97Þ + d3 xdy D0 ðΦ, x, x 0 Þ ½mðxÞ,
^ ^
½ϱ, mðyÞ ,
∂t ∂pi δπ Φ ðxÞ 2

where Hc is the purely classical gravity Hamiltonian. We have


where the Hamiltonian and momentum constraints tell us that
taken the dynamics, i.e., the drift to be local in x, while we allow
ρ(Φ, πΦ, xi, pi) should only have support over phase space degrees of
for the decoherence and diffusion terms to have some range. In
freedom which satisfy the constraint πΦ ρ(Φ, πΦ) = 0. From Eq. (97) we
this case, the evolution law is still local but correlations can be
can identify the classical drift associated to the back-reaction of the
created122. One can also add extra diffusion and decoherence into
matter on the gravitational field from the mðxÞ δπδρðxÞ term, so that
Φ Eq. (101) which we do not consider here since it only leads to
worse experimental bounds. However, adding additional diffu-
Dbr
1, π Φ ðxÞ =  mðxÞ: ð98Þ sion in Φ will generally be required in order to impose the con-
straint πΦ ≈ 081. This master equation is close to the one
In the classical-quantum case, we promote m(x) to an operator m.^
considered in ref. 24, where the decoherence and diffusion ker-
In this case Eq. (93) is the interaction Hamiltonian used in ref. 24 to nels are chosen to be the ones discussed in the second example
study CQ gravity. We see from Eq. (97) that in any theory whose first of Examples of Kernels saturating the decoherence diffusion
moment reproduces the Newtonian back-reaction on average coupling constants trade-off. This is the weak field limit of the
Z   simplest realisation in ref. 11. The case where the diffusion is
  δρ
Tr fH I , ϱg =
3
^
d x Tr mðxÞ ð99Þ spatially uncorrelated D2 ðx, yÞ = ϵðx  x 0 Þ a regulator which
δπ Φ ðxÞ approaches a scalar delta function correspondsp toffiffiffiffiffiffiffiffiffi
the Newtonian
pffiffiffiffiffiffiffiffiffiffi
limit of the diffusion term ϵðx  x 0 ÞfNðxÞ gðxÞ, f gðx 0 Þ, ϱgg.
must have a Dbr
1, π Φ given by Another natural diffusion kernel is
D2 ðx, yÞ =  D2 ð1 + Φðx 0 ÞÞΔx0 δðx, yÞ, which can be understood as the
hDbr ^
1π Φ ðxÞi =  hmðxÞi, ð100Þ Newtonian limit of the spatial diffeomorphism invariant kernel
discussed in the section “Diffeomorphism invariant kernel”.
from which the discussion at the beginning of the section “Physical One can supplement the kernel by some mechanism to control
constraints on the classicality of gravity” follows. the diffusion. For example, a friction term such as
Note that the present discussion is insensitive to the details of the
R pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi
theory provided it satisfies Poisson’s equation on average. None- F ðϱÞ = Df 1
2 dxdx 0 dyfNðxÞ gðxÞ, f gðx 0 Þϵðx  x 0 Þ, HðyÞgϱg: ð102Þ
theless, it’s interesting that when starting from the relativistic theories
of refs. 11,39, we find that the weak field limit resembles the models of In the weak field limit, this would add a term proportional to
Tilloy and Di´osi33, which is discontinuous in Φ, rather than the con-
Z
tinuous model of Di´osi24. This is because24 allows for non-zero con- δ  
F ðϱÞ ≈ dxdx 0 π ðx 0 Þϱ ð103Þ
jugate momentum πΦ with the kinetic energy of a different sign, while δπ Φ ðxÞ Φ
in ref. 81, the momentum is set to zero via the constraint equations and
gauge fixing of the lapse. The discontinuity then arises because we are to the master equation of Eq. (101). Such a term would break Lorentz
operating in the c → ∞ limit, while in ref. 33, the discontinuity arises due invariance since it sets a temperature scale, although this is not
to sourcing the Newtonian potential via a weak measurement process. necessarily a deal breaker, since it is believed by many that quantum
We refer the reader to ref. 81 for details. gravity is also likely to also have an anomaly. However, the friction
term is a modification to D1(x), and if too large, could run afoul of
Weak field CQ master equations precision tests of General Relativity, such as the orbital decay of binary
Although the trade-off we derive does not depend on the parti- pulsars.
culars of the classical–quantum theory (provided it reproduces
Newtonian gravity in the classical limit), we give two concrete Discrete master equation
examples for completeness. In ref. 30 we show that there are two An example of a discrete master equation satisfying Eq. (100) is
classes of classical-quantum dynamics, one which is continuous in
Z
phase space, and one which has discrete jumps in phase space. ∂ϱ c2
≈ fH c ðΦÞ, ϱg  i½H 0ðmÞ , ϱ + d3 x
We will give examples of each. Although they are the weak field ∂t _τ
" R     # ð104Þ
limit of Oppenheim11, it is worth stressing that taking the New- _τ
dyϵðxyÞ 1 + 2ΦðyÞ δ
δπ Φ ðyÞ 2ΦðxÞ y 1
ec2 c2
1 ψðxÞϱψ ðxÞ  fmðxÞ, ϱg + ,
tonian limit entails certain coordinate choices and restrictions on c2 2
the metric. For example, here, we have restricted ourselves to
metrics of the form of Eq. (89). Any gauge fixing of General ^
with τ a dimensionless constant, and mðxÞ = ψy ψ. We have here inclu-
Relativity which is done before deriving the master equation, is ded ℏ and c to make it easier to compare with experiments. To

Nature Communications | (2023)14:7910 15


Article https://doi.org/10.1038/s41467-023-43348-2

leading order, we could drop terms proportional to Φ(x)/c2 in both the In total then, we arrive at the expression for the D2 which saturates
pffiffiffi
exponential and in N g ≈1  2Φ=c2 inside the integral over x. This the bound
gives
1 br μα m2 μα
Z  R  D2 ðx, yÞ = ðD Þ ðxÞ 3 0 Fðx, yÞg N ðx, yÞðDbr*
∂ϱ c2 1 Þi ðyÞ: ð110Þ
≈ fH c ðΦÞ, ϱg  i½H ðmÞ
3 _τ
dyϵðxyÞδπ δ ðyÞ y 1 2 1 i r0λ
0 , ϱ + d x e c2 Φ ψðxÞϱψ ðxÞ  fmðxÞ, ϱg + :
∂t _τ 2
ð105Þ If we further take the back-reaction that of the Newtonian limit in
the section “Newtonian limit of CQ theory”
μα
1 0m π Φ
These dynamical equations are supplemented with modified ðDbr
1 iÞ ðx, yÞ = 2 δ δ i δðx, yÞ then we find the D2 which saturates the
constraint equations as outlined in ref. 37. In any case, the trade-off in bound is
Eq. (31) is a statement independent of constraints and constraint pre-
servation, at least in the weak field limit. 1 m20
D2 ðx, yÞ = Fðx, yÞg N ðx, yÞ: ð111Þ
8 r 30 λ
Examples of Kernels saturating the decoherence diffusion cou-
pling constants trade-off Another example of a Lindbladian coupling which is familiar in the
In this section, we give examples of kernels satisfying the decoherence literature is,
diffusion coupling constant trade-off in Eq. (23). For any choice of
kernel, we can compute the degree of diffusion it induces in precision Dαβ
mass measurements (see the section “Detecting gravitational diffu- Dαβ
0 ðx, yÞ =
0
: ð112Þ
jx  yj
sion”) and decoherence experiments (see the section “Decoherence
rates”) which allows us to rule out certain kernels experimentally. For a single Lindblad operator, this is the coupling introduced in24
Diffeomorphism invariance may single out having both D0(x, y) and used to reproduce a CQ master equation of gravity with a decoherence
D2(x, y) approach delta functions at short distances39, but other alter- rate given by the Diosi–Penrose formula128–130. Here we consider the
natives may be possible. special case where the x, y dependence of D0(x, y) is the same for all
As a first example, we shall take the Lindbladian coupling to be α, β which need not hold in general. The fact that it gives the same
Gaussian, taking decoherence rate as Diosi–Penrose can be seen by plugging Eq. (112)
into the classical-quantum master equation in Eq. (101).
λαβ r 30 To invert the kernel in Eq. (112) we use the fact that
Dαβ
0 ðx, yÞ = g ðx, yÞ ð106Þ
m20 N  
1 2 1
where g N ðx, yÞ is a normalised Gaussian distribution. The mass m0 is a  ∇x = δðx, yÞ, ð113Þ
4π jx  yj
reference mass, and we shall take it equal to the mass of the nucleons
which were considered in the section “Physical constraints on the from which one can immediately read of the generalised inverse
classicality of gravity”, meanwhile λαβ is a coupling constant that ðD1
0 Þαβ ðx, yÞ to be
determines the strength of the Lindbladian.
It should be noted that with this choice of smearing function,
ðD1
0 Þαβ
ðD1 ∇2y ðδðx, yÞÞ, ð114Þ
0 Þαβ ðx, yÞ =
the pure Lindbladian evolution appearing in Eq. (101) can be taken

to resemble the Lindbladian part of spontaneous collapse
models60,99,123–125, except here, there is no need to think about any ad- where ðD1
0 Þαβ are the matrix elements of the generalised inverse of D0.
hoc field, nor think of the collapse as being a physical process. Rather, As a consequence, we find for this specific choice of kernel that the
one necessarily gets decoherence of the wave function for free, via diffusion matrix saturating the coupling constants bound in Eq. (31) is
gravitationally induced decoherence11,25,33,35,126
We now find the diffusion kernel D2(x, y) using the coupling 1
1 μα ðD0 Þαβ 2 ð115Þ
constants trade-off in (23). For simplicity, we shall assume the trade-off Dμν
2, ij ðx, yÞ = D1, i ðxÞ ∇y ðδðx, yÞÞDβν
1, j ðyÞ,
is saturated, and we will take the back-reaction to be local, so that 2 4π
μα br μα
ðDbr
1 Þi ðx, yÞ = ðD1 Þi ðxÞδðx, yÞ. In this case we find where we have also assumed the back-reaction is local. Taking the
back-reaction to further be that of the Newtonian limit of Eq. (101)
1 br μα m2 μα μα
1 0m π Φ
Dμν
2, ij ðx, yÞ = ðD Þ ðxÞ 3 0 g 1 ðx, yÞðDbr*
1 Þi ðyÞ, ð107Þ ðDbr
1 Þi ðxÞ = 2 δ δ i we find
2 1 i r0λ N
1 ðD1
0 Þ 2 ð116Þ
where g 1 N ðx, yÞ is the kernel inverse of a normalised Gaussian D2 ðx, yÞ = ∇ ðδðx, yÞÞ:
distribution. 8 4π y
It is shown in ref. 127, that the inverse distribution takes the form
This diffusion kernel is argued for on the grounds of having the
fluctuations satisfy a Poisson equation, in ref. 24.
g 1
N ðx, yÞ = Fðx, yÞg N ðx, yÞ, ð108Þ
Diffeomorphism invariant kernel
where
Attempts to derive the constraint algebra of a generally covariant CQ
  theory37 (and J. Oppenheim, “The constraints of a continuous realisa-
d X
Y N
x  yi
Fðx, yÞ = cn ðr 0 ÞH 2n i , ð109Þ tion of post-quantum-classical gravity", manuscript in preparation),
i=1 n=0
r0 motivates the spatial diffeomorphism invariant kernel
ð1Þn ðr 0 Þ2n n!
and the limit N → ∞ is taken. In Eq. (109) cn ðr 0 Þ = and d is the
2n
0 1 pffiffiffiffiffiffiffiffiffi
spatial dimension, so that x = (x1, x2, …, xd). Dijkl
2 ðx, x Þ =  D gðxÞNðxÞg ij g kl Δx0 δðx, x 0 Þ, ð117Þ
8

Nature Communications | (2023)14:7910 16


Article https://doi.org/10.1038/s41467-023-43348-2

where Δx is the Laplace–Beltrami operator. One can also consider the However, it is worth noting that while D2 ðx, x 0 Þ may appear
full 3 + 1 kernel, via Δð4Þ δðx, x 0 Þδðt, t 0 Þ along with the associated Green’s to diverge at a R single point as x ! x 0 , when integrated over
function of Δ(4) but this is irrelevant for the Newtonian limit. It is test functions, dxdx 0 D2 ðx, x 0 Þf ðxÞf ðx 0 Þ is usually well-behaved.
however useful in removing the apparent asymmetry in the expres- The kernels discussed above have this property. When it comes
sions below, since one must recall that the δðx, x 0 Þ is a scalar in the first to physically relevant quantities, such as measuring the gravita-
coordinate and a tensor density in the second, and likewise δðt, t 0 Þ has tional diffusion in tabletop experiments, it is the smeared well-
an implicit lapse Nðx 0 Þ in the second position. This kernel’s weak field behaved quantity that is physically relevant. However, in cos-
limit is mology, we typically take the constraint equation of General
Relativity to be exactly satisfied at each point, and so one
Dijkl 0 ij kl 0 ð118Þ might imagine that π 2Φ ðxÞ, and hence D2(x, x) is the relevant
2 ðx, x Þ =  8 Dδ δ ð1 + ΦðxÞÞΔx 0 δðx, x Þ,
1
quantity (see the discussion in the section “Detecting gravita-
which is close to that of Eq. (116), but with a correction term that turns tional diffusion”). However, one can set πΦ ≈ 0 via a gauge
out to be important. freedom81. In GR, the counterpart to the π 2Φ kinetic term is
pffiffiffi Gijklπijπkl(x), and it’s perhaps worth noting that this quantity is not
Using D1 ðx, x 0 Þ =  21 N g δðx, x 0 Þ, the Lindbladian kernel in
dimension d which saturates the trade-off for this diffusion kernel is positive definite. It is also important to note that here, we have
taken the weak field and c → ∞ limit of General Relativity. At short
1 pffiffiffiffiffiffiffiffiffi distances when the diffusion becomes large, we expect this
D0, ijkl ðx, x 0 Þ = 2
gðxÞNðx 0 Þg ij ðxÞg jk ðx 0 ÞGðx, x 0 Þ, ð119Þ approximation to break down. One possible method of studying
2d D
this problem rigorously would be through the regularisation
properties of the classical-quantum path integral which we
with Gðx, x 0 Þ the Green’s function for −Δ. It is a density in the x 0 coor-
introduce in38,39.
dinate and a scalar in x. In the weak field limit, and to 0th order in Φ(x),
this gives the Diosi–Penrose kernel, Eq. (112).
Decoherence rates
One could also consider the kernel
In this section, Rwe relate decoherence rates to D0, and also to the
average hD0 i = dzTr½Dαβ y
0 ðz; x, yÞLα ðxÞϱLβ ðyÞ. In particular, we shall
0 1 pffiffiffiffiffiffiffiffiffi ij kl
Dijkl
2 ðx, x Þ =  D gðxÞg g Δx NðxÞδðx, x 0 Þ, ð120Þ show that the decoherence rate of a mass in superposition, is given by
8
Eq. (131) in terms of the Lindblad operators and Dαβ 0 , and can be related
to the quantity 〈D0〉.
which in the weak field limit is We consider the case of a quantum mass initially in a partially
decohered superposition of state jLi and jRi. We describe the
Dijkl 0 1 ij kl0
2 ðx, x Þ =  8 Dδ δ ΔΦðxÞδðx, x Þ:
ð121Þ quantum state using creation and annihilation operators
ψ(x), ψ†(x) on a Fock space, related to the usual momentum-based
R
Fock operators as ψðxÞ = dpei~p~x a~p . The mass density operator is
A comment on divergences ^
defined via mðxÞ = mψy ðxÞψðxÞ, where m is the mass of the particle.
The kernel examples given above give rise to divergent variance in the We assume that the state remains well approximated by a state
classical degrees of freedom, since in both cases the diffusion coeffi- with fixed particle number, and the superposition can be taken to
cient diverges when evaluated at the same point D2(x, x). Though we be distributions centred around x = xL and x = xR with total mass
do not have a general proof, this seems to be a general feature of the M, i.e., for a one-particle state we could take
coupling constant trade-off: for the examples where we can compute R 3 y
jL=Ri = d xf L=R ðxÞψ ðxÞj0i. We will take them to be well sepa-
the kernel inverse, at least one of D2(x, x) and D0(x, x) diverge. A
divergent D2(x, x) generally leads to a formally divergent classical rated so that fL(x)fR(x) ≈ 0, and we take the separation distance to
energy production, whilst a divergent Lindbladian coupling D0(x, x) be larger than the scale of the non-locality in D0(x, y). Mathema-
can lead to a divergent energy production in the matter degrees of tically this means that hLjDαβ y
0 ðz; x, yÞLβ ðyÞLα ðxÞjRi ≈ 0 for any local
freedom. The latter is related to the BPS problem98 of anomalous operators Lα(x) and Lβ(y).
heating, although it isn’t necessarily equivalent since some kernels may
diverge and be well-behaved from the point of view of energy pro- With this orthogonality condition, we can then (at least initially)
duction. This is not an issue from a conceptual point of view, since the consider the joint quantum classical state restricted to the two-
only reason we expect energy to be conserved is due to Noether’s dimensional Hilbert space of these two states so that the total
theorem, and Noether’s theorem doesn’t apply when the evolution quantum-classical system can be written as
isn’t unitary.
 
In the standard BPS problem, energy production in open uL ðΦ, π Φ , tÞ αðΦ, π Φ , tÞ
ϱðΦ, π Φ , tÞ = , ð122Þ
quantum field theory can be made small by renormalizing the ?
α ðΦ, π Φ , tÞ uR ðΦ, π Φ , tÞ
Lindbladian coefficient D0(x, y) appearing in the master equation.
Thus the problem is merely one akin to the hierarchy problem, where uL(Φ, πΦ, t) and uR(Φ, πΦ, t) correspond to some subnormalised
where we are required to introduce another energy scale. How- probability distribution over the classical states of the
ever, in the case of classical–quantum coupling, the coupling gravitational field.
constant trade-off tells us that we cannot re-normalise D0(x, y) We define the total quantum state ρQ by integrating over the
without affecting D2(x, y). In particular, tuning the diagonals classical degrees of freedom
D0(x, x) to be arbitrarily small (large) has the effect of tuning Z
D2(x, x) to be arbitrarily (large) small: heuristically, one trades
ρQ = DΦDπ Φ ϱðΦ, π Φ , tÞ, ð123Þ
energy production in the classical system with energy production
in the quantum system, and the relationship is fixed by the trade-
off. On expectation, the total energy could be preserved, and the and we shall relate 〈D0〉 appearing in the trade-off to the
back-reaction can even slow down the flow of energy, but it’s decoherence rate of the off diagonals of ρQ. Integrating over
unclear if this is enough. the classical phase space in Eq. (9), one finds the following expression

Nature Communications | (2023)14:7910 17


Article https://doi.org/10.1038/s41467-023-43348-2

for the evolution of ρQ Equation (129) already expresses the fact that the off-diagonal
terms will decay, and the particle will decohere at a rate determined by
Z
∂ρQ the integrand of Eq. (129).
= DϕDπ Φ  i½HðΦ, π Φ Þ, ϱðΦ, π Φ Þ
∂t We can go slightly further when in the presence of a background
Z Z h Newtonian potential which is dominant, such as the Earth’s Φb. The
+ DϕDπ Φ dxdy Dαβ y
0 ðΦ, π Φ ; x, yÞLα ðxÞϱðΦ, π Φ , tÞLβ ðyÞ Earth’s background potential dominates over small fluctuations in Φ

1 due to the particles and we can approximate Eq. (129) by
 Dαβ ðΦ, π Φ ; x, yÞfL y
β
ðyÞL α ðxÞ, ϱðΦ, π Φ , tÞg :
2 0
ð124Þ 1
 Dαβ ðx, yÞðhLjLyβ ðyÞLyβ ðyÞLα ðxÞjLi + hRjLyβ ðyÞLyβ ðyÞLα ðxÞjRiÞhLjρQ jRi,
2 0
∂ρ
In particular, one finds that the off-diagonals hLj ∂tQ jRi evolve in ð130Þ
part according to the commutator, and in part due to the Lindbladian
term where the coupling Dαβ 0 ðx, yÞ depends on the background Newtonian
Z Z h potential, but is otherwise phase-space independent. The result is to
DϕDπ Φ dxdy hLjDαβ y
0 ðΦ, π Φ ; x, yÞLα ðxÞϱðΦ, π Φ , tÞLβ ðyÞjRi exponentially decrease the coherence 〈L∣ρQ∣R〉 with a rate λ deter-
 ð125Þ mined by
1
 Dαβ y
0 ðΦ, π Φ ; x, yÞhLjfLβ ðyÞLα ðxÞ, ϱðΦ, π Φ , tÞgjRi :
2 Z
1
λ= dxdyDαβ y y
0 ðx, yÞðhLjLβ ðyÞLα ðxÞjLi + hRjLβ ðyÞLα ðxÞjRiÞ:
ð131Þ
2
Care must be taken however, because both the quantum Hamil-
tonian and the Lindbladian coupling constants depend on the classical
degrees of freedom which are affected by the quantum degrees of Let us now show that the 〈D0〉 term appearing in the trade-off (34)
freedom, and thus the evolution of the quantum system is non- is always less than (twice) this decoherence rate when in the presence
Markovian in general. of a background potential. Specifically, we show that
We shall now study the two terms appearing in Eq. (125) sepa- Z Z h i
rately, starting with the first term. Since we assume that the state is hD0 i = dxdyTr Dαβ y
DΦDπ Φ 0 ðΦ, π Φ ; x, yÞLβ ðyÞLα ðxÞϱðΦ, π Φ Þ ≤ 2λ,
well approximated by a state with fixed particle number then the
contributions to the first term in Eq. (125) only come from ð132Þ
terms where Lα(x) and Lβ(y) have the same number of creation and
annihilation operators. To compute the expression, one where we assume that we are in the presence of a background
commutes through the creation operators to act on the hLj bra, and potential. To see this, we first expand out the CQ state in terms of
picks up a term fL(x). Similarly, one commutes the annihilation Eq. (128) and use the fact that D0 has a range less than the separation of
operators to the act on the jRi ket, and picks up a term fR(y). As a the masses. We then arrive at the following expression for the left-hand
consequence side of Eq. (132)
Z Z
hLjDαβ y αβ
0 ðΦ, π Φ ; x, yÞLα ðxÞϱðΦ, π Φ , tÞLβ ðyÞjRi ∼ D0 ðΦ, π Φ ; x, yÞf L ðxÞf R ðyÞ ≈ 0, DΦDπ Φ dxdyDαβ y y
0 ðΦ, π Φ ; x, yÞðhLjLβ ðyÞLα ðxÞjLiuL ðΦ, π Φ , tÞ + hRjLβ ðyÞLα ðxÞjRiuR ðΦ, π Φ , tÞÞ,
ð126Þ
ð133Þ
where the last equality follows from the fact that we are taking the
In the presence of a background potential, this dominates the
masses to be well separated and the range of D0(x, y) is assumed to be
contribution to the decoherence and we are left with
much less than the separation between the masses.
Hence, the evolution of the off-diagonals comes from the Z
(off-diagonals) of the unitary evolution and the second term in Eq. dxdyDαβ y y
0 ðx, yÞðhLjLβ ðyÞLα ðxÞjLihLjρQ jLi + hRjLβ ðyÞLα ðxÞjRihRjρQ jRiÞ:
(125), the so-called no-event term. The off-diagonals of the no-event
term is ð134Þ

Z Z
1 Due to the positivity of the CQ density matrix 〈L∣ρQ∣L〉 and 〈R∣ρQ∣R〉
 DϕDπ Φ dxdyDαβ y
0 ðΦ, π Φ ; x, yÞhLjfLβ ðyÞLα ðxÞ, ϱðΦ, π Φ , tÞgjRi,
2 must both be positive. Furthermore, they must sum to one due to
ð127Þ normalisation, from which Eq. (131) directly follows.
It is also important to note that though λ is the decoherence rate
which is negative definite and acts to exponentially suppress the of a particle in superposition of L/R states, the bound (132) holds even
coherence. To see this, note that expanding out ϱ(Φ, πΦ, t) in terms of for fully decohered masses in any mixture of jLihLj, jRihRj states. This
the approximate 2 dimensional Hilbert space can be seen directly from (133) which depends only on uL, uR.

ϱðΦ, π Φ , tÞ = uL ðΦ, π Φ , tÞjLihLj + uR ðΦ, π Φ , tÞjRihRj Decoherence rate examples


ð128Þ
In this section, we give an explicit example of a decoherence rate
+ αðΦ, π Φ , tÞjLihRj + α * ðΦ, π Φ , tÞjRihLj,
calculation. Importantly, we see that in general, the decoherence rate
and using the fact that the range of D0(x, y) is much less than the can depend on the probability density. This suggests that the terms
separation between the left and right masses, we can write the off- appearing in the trade-off relation will need to depend on expectation
diagonals of the no-event term as values such as the expectation value of the mass at a point x, rather
than a stronger bound in terms of the mass density. This is perhaps not
Z  
1 surprising, since the decoherence rate itself can be thought of as an
 DΦDπDαβ y y
0 ðΦ, π Φ ; x, yÞ hLjLβ ðyÞLα ðxÞjLi + hRjLβ ðyÞLα ðxÞjRi hLjϱðΦ, π Φ ÞjRi:
2 expectation value, being related to the average time it takes for off-
ð129Þ diagonal elements to decay. In conclusion, this motivates us to

Nature Communications | (2023)14:7910 18


Article https://doi.org/10.1038/s41467-023-43348-2

advocate for the volume of the wave packet to be included in the figure (140), which gives an extra factor of N relative to the single particle
2
of merit in future interference experiments. case λ = 6D05RNM .
We take the Newtonian limit master equation defined by Eq. (101). Both Eqs. (138) and (139) depend on the probability density of the
We ignore the unitary part of the evolution, since it will not directly mass. In particular, taking the probability density to be arbitrarily
contribute to the decoherence rate, and can be small for a free particle peaked, one finds that the decoherence rate also diverges. This has to
in superposition. From Eq. (101) we find the relevant evolution for the be the case: recall from the section “A Trade-off between decoherence
quantum state ρQ, obtained by integrating over the classical degrees of and diffusion” that if one considers a particle in a superposition of two
freedom to be arbitrarily peaked probability densities, then there can be an arbitrarily
large response in the Newtonian potential around those points. As a
Z
∂ρQ 1   consequence, for such states, the decoherence must occur arbitrarily
= ^
d3 xd3 yD0 ðx, yÞ ½mðxÞ, ^
½ρQ , mðyÞ : ð135Þ
∂t 2 fast, or there must be an arbitrarily large amount of diffusion to cover
up the back-reaction and maintain coherence. For the continuous
We now compute the off-diagonal elements for a particle in master equation, such as that of Eq. (101) this diffusion must also occur
superposition of orthogonal jLi, jRi states throughout space, although it can depend on the gravitational degrees
of freedom. Since divergent energy production throughout space is
  Z
∂ρ  clearly unphysical, it must be the case that the decoherence rate must
L Q R =  d3 xd3 yD0 ðx, yÞ ðmL ðxÞ  mR ðxÞÞðmL ðyÞ  mR ðyÞÞhLjρQ jRi:
∂t also depend on the expected mass density, as is the case for this
ð136Þ example. This argument allows us to rule out continuous master
equations that have pure Lindbladian terms that predict decoherence
^
where mL ðxÞ ≈ hLjmðxÞjLi and similarly for the right state. We see that rates which remain finite as the mass density becomes arbitrarily
the off-diagonals decay exponentially with a rate determined by peaked since the coupling constant trade-off will demand that an
Z infinite amount of diffusion is required to cover up the back-reaction
and maintain coherence. This is the case for the class of models with
λ= d3 xd3 yD0 ðx, yÞðmL ðxÞ  mR ðxÞÞðmL ðyÞ  mR ðyÞÞ: ð137Þ
CSL-type Lindbladian couplings which are phase space independent,
for example in Eq. (106).
In the main body, and the previous subsection, we have assumed
that the superposition of the particle is much less than the typical scale Detecting gravitational diffusion
of D0(x, y). In this example, this means that we take the particles suf- In this section, we show how the diffusion induced on the Newtonian
ficiently separated so that we can approximate D0(x, y)mL(x)mR(y) ≈ 0, potential can be measured experimentally. metric leads to observable
in which case Eq. (137) is precisely the decoherence rate calculated in effects, such as variations in the accelerations involved in torsion
^
Eq. (131) with LðxÞ = mðxÞ, as is to be expected. experiments and stochastic wave production in cosmology. As shown
The natural decoherence kernel from the point of view of dif- in the main body of the text, in the non-relativistic limit, c → ∞, the CQ
feomorphism invariance is D0(x, y) = D0δ(x, y), in which case for a dynamics can be approximated by sourcing the Newtonian potential
particle of mass M, and uniform wave-packet volume V, Eq. (137) gives by a random mass term, and in order to maintain the coherence of any
λ = 2D0M2/V. mass is superposition, there must be noise in the Newtonian potential
As another example, we can take D0 ðx, x 0 Þ to be the Diosi–Penrose such that we cannot tell which element of the superposition the par-
D0
decoherence kernel defined via D0 ðx, yÞ = jxyj , so that the off- ticle will be in
diagonals decay exponentially with a rate proportional to the
Diosi–Penrose decoherence rate ^ Jðx, tÞ,
∇2 Φ = 4πG½mðx, tÞ + uðΦ, mÞ ð141Þ
Z
D0 with
λ= d3 xd3 y ðm ðxÞ  mR ðxÞÞðmL ðyÞ  mR ðyÞÞ: ð138Þ
jx  yj L
Em ½ Jðx, tÞ = 0, Em ½uJðx, tÞuJðy, t 0 Þ = 2hD2 ðx, y, ΦÞiδðt, t 0 Þ, ð142Þ
In this example, taking the superposition to be sufficiently sepa-
rated means that we are approximating jx Dx0
≈ 0 in comparison with where hD2 ðx, y, ΦÞi :¼ Tr½Dμν y
2 ðx, y, Φb ÞLμ ðxÞρLν ðyÞ and ρ is the quan-
L Rj
the rest of the terms appearing in Eq. (138). We are then left with tum state for the decohered mass density. The diffusion coefficient in
Z Eq. (142) is chosen in order for the dynamics to have the same
3 3 D0 moments as the CQ master equation (4). The solution to Eq. (141),
λ= d xd y ðm ðxÞmL ðyÞ + mR ðxÞmR ðyÞÞ, ð139Þ
jx  yj L having absorbed u into J is given by
Z
which for spherical distributions of radius R and total mass M is pro- ^
½mðx 0 , tÞ  uðΦ, mÞJðx 0
, tÞ
d x0
3
Φðt, xÞ =  G , ð143Þ
portional to the average gravitational self-energy of each mass
2
jx  x j
0
0M
distribution λ = 6D5R .
For a composite particle of mass M, made up of N constituents where the statistics of J are described by Eq. (142). A formal treatment
each of radius R, the mass density will be represented by a sum over all of solutions to non-linear stochastic integrals of the form Eq. (141) can
of the particles m(x) = ∑imi(x). The decoherence rate is given by be found in ref. 82.
! One can also verify this behaviour in specific cases. In the con-
Z X X tinuous model of the section “Continuous master equation”, the noise
3 D0
3
λ= d xd y m ðxÞmL, j ðyÞ + mR, i ðxÞmR, j ðyÞ : is taken to be Gaussian, and this, as well as the evolution of the
jx  yj i, j L, i i, j
quantum state, is what determines the diffusion in Eq. (143). For the
ð140Þ class of discrete models, the higher order moments such as
Em ½Jðx, tÞJðy, t 0 ÞJðz, t 00 Þ are suppressed by an order parameter11,26,37 and
Since the cross terms involving i, j are suppressed by a factor of whenever this is true we expect we can approximate the dynamics of
inter-atomic scales, to leading order the contribution to the deco- the Newtonian potential by a Gaussian process. Whether this is the
herence rate is lower bounded by the i = j component of the sums in Eq. case or not, it is the second-order moment that enters into our

Nature Communications | (2023)14:7910 19


Article https://doi.org/10.1038/s41467-023-43348-2

discussion of the variance here. As such, for minimally coupled the- This is true for the kernels with functional dependence of the form
ories, the Newtonian potential will appear to be sourced by a random D2 ~ Φn, D2 ~ ∇ Φ, though the approximation does not hold for all ker-
mass distribution. nels, for example, D2 ~ ∇2Φ which creates diffusion only where there is
In the discrete case, a precise understanding of the effects of the the mass density. We hereby shall only consider diffusion kernels
diffusion beyond the Gaussian approximation involves solving the full D2(x, y, Φb) where the background potential is dominant, leaving more
classical-quantum dynamics, perhaps using the methods of Oppen- general considerations for future work.
heim et al.26. In Eq. (141) we are also taking the time scale of the dif- For local translation invariant dynamics for which the background
fusion to be faster than the dynamics of the matter distribution. Newtonian potential is dominant, for example, D2 ~ Φn, we have
Likewise for the decoherence—we showed in the “Methods” subsection 〈D2(x, y, Φb)〉 = 〈D2(Φb)〉δ(x, y) and we arrive at the expression for the
“Decoherence rates” for continuous models the evolution of the total variation in time-averaged force
quantum state acts to decohere it into a mass density eigenbasis m(x).
Z ! !0 ! !0
One could of course also include the quantum state evolution in a 2G2 X ðx  x Þð y  x Þ
simulation of full CQ dynamics, but this is beyond the scope of the σ 2F = d3 xd3 yd3 x 0 mi ðxÞmj ðyÞ hD2 ðx 0 , Φb Þi:
ΔT ij jx  x 0 j3 jy  x 0 j3
current work.
Note that if the moments of the noise process are Galilean ð147Þ
invariant, then the theory given by Eq. (141) is Galilean invariant as we
would expect in the Newtonian limit. One can further derive Eq. (141) To leading order, the integral in Eq. (147) is dominated by the self
from manifestly diffeomorphism invariant theories81, such as those of variation term where i = j, since nuclear scales 10−15 m dominate over
P P
refs. 39,131. inter-atomic scales 10−9 m, so that E½ ij F i F j  ∼ i E½F 2i . Approximat-
ing the mass density of the atoms as coming from their nucleus, and
Table-top experiments taking them to be spheres of constant density ρ with radius rN and mass
In this section we estimate the variation in force which would be seen mN, we find that the integral in Eq. (147) is approximately
in table-top experiments which bounds the diffusion of classical the-
Z
ories of gravity from above, giving a squeezed bound on D2 due to NG2 ρ2 r 2N
σ 2F ∼ d3 x 0 hD2 ðΦb Þi: ð148Þ
lower bounds on diffusion arising from coherence experiments. We do ΔT
this for dynamics in Eq. (141), but the methodology is general and
could also be used in a full simulation of CQ dynamics. For the class of continuous dynamics 〈D2(Φb)〉 = D2(Φb), since the
The variation in force induced on a composite mass is found via diffusion is not associated with any Lindblad operators. If there is noise
Z everywhere throughout space, then the integral in Eq. (148) diverges
! and gives evidence that continuous CQ theories with noise everywhere
F tot =  d3 xmðxÞ∇Φ: ð144Þ
should be ruled out.
As such, we expect that continuous CQ theory must contain non-
Using the solution in Eq. (143), the total force can be written linear terms proportional to the Newtonian potential appearing in Eq.
R
(141), in which case we can approximate dx 0 D2 by VbD2 where Vb is
Z ! !0 the volume of the region over which the background Newtonian
! ðx  x Þ
F tot =  G d3 xd3 x 0 mðxÞ ½mðx 0 , tÞ  Jðx 0 , tÞ: ð145Þ
jx  x 0 j3 potential is significant. In total then, we find for continuous local CQ
dynamics
In reality, we measure time-averaged force by measuring time-
averaged
R ΔT accelerations over the time resolution of the experiment D2 NG2 ρ2 r 2N V b ð149Þ
σ 2F ∼ :
ΔT ΔT
1
0 dtF tot . The total variation in the force’s time-averaged mag- ΔT
nitude. The full covariance matrix for various kernels in the Newtonian
limit is given in (J. Oppenheim and A. Russo, manuscript in prepara- From this, we can calculate D2 in terms of the total variance of the
! ! σ2
tion), σ 2F :¼ F tot  F tot can be written as acceleration σ 2a = m2F to get a lower bound
tot

Z ! !0 ! !0
1 ðx  x Þð y  y Þ σ 2a Nr 4N ΔT
σ 2F = 2G2 d xd yd x 0 d y0 mðxÞmðyÞ
3 3 3 3
hD2 ðx 0 , y0 , ΦÞi: D2 ≤ : ð150Þ
ΔT jx  x 0 j3 jy  y0 j3 V b G2
ð146Þ
Standard Cavendish type classical torsion experiments measure
We shall use Eq. (146) to provide an upper bound on coupling accelerations of the order 10−7 m s−2, and we can take the time over
constants of CQ theories for different choices of kernels D2 ðx 0 , y0 , ΦÞ. which the acceleration is averaged to be that of minutes ΔT ~ 102 s, so a
Given a choice of functional form of the kernel, all that remains is the very conservative bound is σa ~ 10−7 m s−2, whilst N will be N ~ 1026 and
strength of the diffusion coupling, which for the translation invariant rN ~ 10−15 m. We take the background Newtonian potential to be that of
kernels we consider here takes the form of a single coupling constant the Earth and we (conservatively) take Vb to be V b ∼ r 2E h ∼ 1015 m3
D2. We take D2 to be a dimension-full quantity with units kg2 sm−3 which where rE is the Earth’s radius and h is the atmospheric height. We see
characterises the rate of diffusion for the conjugate momenta of the that this bounds D2 from above by D2 ≤ 10−41 kg2 sm−3.
Newtonian potential. On the other hand, D2 is bounded from below from interferometry
For a composite mass, we can approximate the mass density by experiments which bound the decoherence rate. From Eq. (137) and the
summing over N individual atoms of mass density mi(x), m(x) = ∑imi(x). coupling constant trade-off, for the kernel D2(x, y) = D2δ(x, y) we see
! P! !
The total force is the given by F tot = i F i , where F i is the force on (ignoring constant factors) that the decoherence rate is found to be
! R
each individual atom F i =  V dxmi ðxÞ∇ΦðxÞ, and the total variation
P P
of force is then σ 2F = E½ ij F i F j   E½ i F i 2 . N λ M 2λ
λ∼ , ð151Þ
In general, the squeeze will depend on the functional choice of V λ D2
D2(x, y, Φ) on the Newtonian potential. As mentioned in the main body,
in the presence of a large background potential Φb, such as that of the where Mλ is the mass of a composite particle in the interferometry
Earth’s, we will often be able to approximate D2(x, y, Φ) = D2(x, y, Φb). experiment, which is made up of Nλ particles, each with volume Vλ. This

Nature Communications | (2023)14:7910 20


Article https://doi.org/10.1038/s41467-023-43348-2

gives rise to the squeeze In general then, we expect that by simulating full CQ dynamics
satisfying the decoherence diffusion trade-off we will be able to
σ 2a Nr 4N ΔT N λ M 2λ squeeze D2 from above and below. We bound D2 from above by
≥ D2 ≥ : ð152Þ studying the effects of diffusion on gravitational experiments, and we
V bG 2 V λλ
bound D2 from below using the coupling constant trade-off and
Using the numbers from ref. 59, with Mλ ~ 10−24 kg, Nλ ~ 103, and coherence experiments lower bounding the decoherence rate. As we
Vλ ~ 10−1510−1510−7 m3 = 10−37 m3, λ ~ 101 s−1 we find that D2 ≥ 10−9 kg2 sm−3. have seen in this section, it appears that classes of continuous CQ
This suggests that the D2(x, y) = D2δ(x, y) kernel for classical gravity is hybrid theories of gravity, including models without spatial correla-
already ruled out by experiment. tions, are already experimentally ruled out, whilst others, such as the
For the local discrete models, such as that of Eq. (104), the theory kernels in subsection “Diffeomorphism invariant kernel” require
is less constrained due to the dependence of the diffusion on the mass stronger bounds from both gravitational and coherence experiments.
3
l We have been very conservative in our estimates, and so we expect a
density. In this case hD2 ðΦb Þi = mP D2 ðΦb ÞmðxÞ, where the factors of
P
Planck length and Planck mass are to ensure that D2(Φb) has the more thourough analysis will tighten the bounds by orders of
required units. We arrive at the upper bound for D2 magnitute.

σ 2a Nr 4N ΔTmP References
3
≥ D2 : ð153Þ 1. DeWitt, C. M. & Rickles, D. The Role of Gravitation in Physics: Report
mN G2 l P
from the 1957 Chapel Hill Conference, Vol. 5 (epubli, 2011).
Meanwhile, from Eq. (131), and coupling constant trade-off (30) 2. Feynman, R. P. In Feynman Lectures on Gravitation (eds Morinigo,
the decoherence rate for local discreet jumping models goes as F. B., Wagner, W. G. & Hatfield, B.) 10–11 (1996).
M λ mP 3. Aharonov, Y. & Rohrlich, D. Quantum Paradoxes: Quantum Theory
λ∼ 3 , which gives rise to the lower bound for D2. From this, we
l P D2 for the Perplexed 212–213 (Wiley-VCH, 2003).
arrive at the squeeze 4. Eppley, K. & Hannah, E. The necessity of quantizing the gravita-
tional field. Found. Phys. 7, 51 (1977).
3
σ 2a Nr 4N ΔT l P D2 M λ 5. Unruh, W. G. Steps towards a quantum theory of gravity. In
≥ ≥ , ð154Þ
mN G 2 mP λ Quantum Theory of Gravity: Essays in honor of the 60th birthday
of Bryce S. DeWitt (ed Christensen, S. M.) (Adam Hilger Ltd., 1984).
6. Carlip, S. Is quantum gravity necessary? Class. Quantum Gravity
and plugging in the numbers we find the bound given by Eq. (46) which
25, 154010 (2008).
gives rise 3to the squeeze for local discrete mod-
7. Mari, A., De Palma, G. & Giovannetti, V. Experiments testing
els 101 kgs ≥ mP D2 ≥ 1025 kgs.
l
P macroscopic quantum superpositions must be slow. Sci. Rep. 6,
We can also consider other diffusion kernels, for example, that of
22777 (2016).
Eq. (118). In this case, for continuous dynamics, we have that
2 8. Baym, G. & Ozawa, T. Two-slit diffraction with highly charged
hD2 ðx, yÞi =  l P D2 ðΦb Þ∇2 δðx, yÞ. The Lindbladian kernel saturating the
particles: Niels Bohr’s consistency argument that the electro-
coupling constants trade-off at zeroeth order in Φ(x), is the
0 ðΦb Þ magnetic field must be quantized. Proc. Natl Acad. Sci. USA 106,
Diosi–Penrose kernel D0 ðx, y, Φb Þ = Djxyj , as we saw in the section
3035 (2009).
“Examples of Kernels saturating the decoherence diffusion coupling
9. Belenchia, A. et al. Quantum superposition of massive objects and
constants trade-off”. Approximating the masses as spheres of constant
the quantization of gravity. Phys. Rev. D 98, (2018) https://doi.org/
density we find from a substitution of the kernel into Eq. (146) that the
10.1103/physrevd.98.126009.
variation in time-averaged force is given by
10. Kent, A. Simple refutation of the Eppleyhannah argument. Class.
2 Quantum Gravity 35, 245008 (2018).
l P G2 m2N ND2
σ 2F ∼ : ð155Þ 11. Oppenheim, J. A post-quantum theory of classical gravity?.
ΔTr 3N arXiv:1811.03116 [hep-th] (2018).
12. Rydving, E., Aurell, E. & Pikovski, I. Do gedanken experiments
We therefore find a lower bound for D2 in terms of the variation in
compel quantization of gravity? Phys. Rev. D 104, 086024
acceleration
(2021).
13. Aleksandrov, I. & Naturf, Z. 36a, 902 (1981); a. anderson. Phys. Rev.
2
ΔTl P σ 2a Nr 3N Lett 74, 621 (1995).
D2 ≤ , ð156Þ
G2 14. Kapral, R. Progress in the theory of mixed quantum-classical
dynamics. Annu. Rev. Phys. Chem. 57, 129–157 (2006).
which for classical torsion experiments σa ~ 10−7 m s−2, T ~ 102 s, N ~ 1026 15. Boucher, W. & Traschen, J. Semiclassical physics and quantum
and rN ~ 10−15 m gives D2 l p ≤ 109 kg m1 . On the other hand, for this
2
fluctuations. Phys. Rev. D 37, 3522 (1988).
kernel the decoherence rate can be calculated via Eq. (139) 16. Diósi, L., Gisin, N. & Strunz, W. T. Quantum approach to
coupling classical and quantum dynamics. Phys. Rev. A 61,
NM 2λ 022108 (2000).
λ∼ 2
, ð157Þ
l P D2 Rλ 17. Sato, I. An attempt to unite the quantum theory of wave field with
the theory of general relativity. Sci. Rep. Tohoku Univ. 1st ser. Phys.
which gives the squeeze on D2 Chem. Astron. 33, 30 (1950).
18. Møller, C. et al. Les théories relativistes de la gravitation. In Col-
ΔTσ 2a Nr 3N N M2 loques Internationaux CNRS Vol. 91 (1962).
≥ λ λ:
2
≥ l P D2 ð158Þ
G2 Rλ λ 19. Rosenfeld, L. On quantization of fields. Nucl. Phys. 40, 353 (1963).
20. Page, D. N. & Geilker, C. Indirect evidence for quantum gravity.
For the numbers used in the main body of the Phys. Rev. Lett. 47, 979 (1981).
text59Mλ ~ 10−24 kg, Nλ ~ 103, Rλ ~ V1/3 = 10−12 m, λ ~ 101 s, this yields 21. Blanchard, P. & Jadczyk, A. Event-enhanced quantum theory and
1
D2 l P ≥ 1035 kg m and so this model is not ruled out by experiment.
2
piecewise deterministic dynamics. Ann. Phys. 507, 583 (1995).

Nature Communications | (2023)14:7910 21


Article https://doi.org/10.1038/s41467-023-43348-2

22. Diosi, L. Quantum dynamics with two Planck constants and the 47. Barceló, C., Carballo-Rubio, R., Garay, L. J. & Gómez-Escalante, R.
semiclassical limit. arXiv:quant-ph/9503023 [quant-ph] (1995). Hybrid classical-quantum formulations ask for hybrid notions.
23. Alicki, R. & Kryszewski, S. Completely positive Bloch–Boltzmann Phys. Rev. A 86, 042120 (2012).
equations. Phys. Rev. A 68, 013809 (2003). 48. Marletto, C. & Vedral, V. Why we need to quantise everything,
24. Diósi, L. The gravity-related decoherence master equation from including gravity. npj Quantum Inf. 3, 29 (2017).
hybrid dynamics. J. Phys.-Conf. Ser. 306, 012006 (2011). 49. Cavendish, H. Xxi experiments to determine the density of the
25. Poulin, D. & Preskill, J. Information loss in quantum field theories. earth. Philos.Trans. R. Soc. Lond. 88, 469-526 (1798).
Front. Quantum Inf. Phys. KITP https://online.kitp.ucsb.edu/ 50. Luther, G. G. & Towler, W. R. Redetermination of the Newtonian
online/qinfo_c17/poulin/ (2017). gravitational constant g. Phys. Rev. Lett. 48, 121 (1982).
26. Oppenheim, J., Sparaciari, C., Šoda, B., & Weller-Davies, Z. 51. Gundlach, J. H. & Merkowitz, S. M. Measurement of Newton’s
Objective trajectories in hybrid classical-quantum dynamics. constant using a torsion balance with angular acceleration feed-
Quantum 7, 891 (2023). back. Phys. Rev. Lett. 85, 2869 (2000).
27. Layton, I., Oppenheim, J. & Weller-Davies, Z. A healthier 52. Quinn, T. Measuring big g. Nature 408, 919 (2000).
semi-classical dynamics. arxiv2208.11722 (2022). 53. Gillies, G. & Unnikrishnan, C. The attracting masses in measure-
28. Gorini, V., Kossakowski, A. & Sudarshan, E. C. G. Completely ments of g: an overview of physical characteristics and perfor-
positive dynamical semigroups of N-level systems. J. Math. Phys. mance. Philos. Trans. R. Soc. A: Math. Phys. Eng. Sci. 372,
17, 821 (1976). 20140022 (2014).
29. Lindblad, G. On the generators of quantum dynamical semi- 54. Rothleitner, C. & Schlamminger, S. Invited review article: mea-
groups. Commun. Math. Phys. 48, 119 (1976). surements of the Newtonian constant of gravitation, g. Rev. Sci.
30. Oppenheim, J., Sparaciari, C., Šoda, B. & Weller-Davies, Z. The two Instrum. 88, 111101 (2017).
classes of hybrid classical–quantum dynamics. arXiv:2203.01332 55. Arndt, M. et al. Wave–particle duality of c60 molecules. Nature
[quant-ph] (2022). 401, 680 (1999).
31. Kafri, D., Taylor, J. M. & Milburn, G. A classical channel model for 56. Nimmrichter, S., Hornberger, K., Haslinger, P. & Arndt, M. Testing
gravitational decoherence. New J. Phys. 16, 065020 (2014). spontaneous localization theories with matter-wave inter-
32. Kafri, D., Milburn, G. J. & Taylor, J. M. Bounds on quantum ferometry. Phys. Rev. A 83, 043621 (2011).
communication via Newtonian gravity. N. J. Phys. 17, 015006 57. Juffmann, T. et al. Real-time single-molecule imaging of quantum
(2015). interference. Nat. Nanotechnol. 7, 297 (2012).
33. Tilloy, A. & Diósi, L. Sourcing semiclassical gravity from sponta- 58. Juffmann, T., Nimmrichter, S., Arndt, M., Gleiter, H. & Hornberger,
neously localized quantum matter. Phys. Rev. D 93, 024026 K. New prospects for de Broglie interferometry. Found. Phys. 42,
(2016). 98 (2012).
34. Tilloy, A. & Diósi, L. On gkls dynamics for local operations and 59. Gerlich, S. et al. Quantum interference of large organic molecules.
classical communication. Open Syst. Inf. Dyn. 24, 1740020 Nat. Commun. 2, 263 (2011).
(2017). 60. Bassi, A. & Ghirardi, G. Dynamical reduction models. Phys. Rep.
35. Tilloy, A. & Diósi, L. Principle of least decoherence for Newtonian 379, 257–426 (2003).
semiclassical gravity. Phys. Rev. D 96, 104045 (2017). 61. Westphal, T., Hepach, H., Pfaff, J. et al. Measurement of gravita-
36. Arnowitt, R., Deser, S. & Misner, C. W. Republication of: the tional coupling between millimetre-sized masses. Nature 591,
dynamics of general relativity. Gen. Relativ. Gravit. 40, 1997 225–228 (2021).
(2008). 62. Schmöle, J., Dragosits, M., Hepach, H. & Aspelmeyer, M. A
37. Oppenheim, J., Weller-Davies, Z. The constraints of post-quantum micromechanical proof-of-principle experiment for measuring the
classical gravity. JHEP. 2022, 80 (2022). gravitational force of milligram masses. Class. Quantum Gravity
38. Oppenheim, J. & Weller-Davies, Z. Path integrals for classical- 33, 125031 (2016).
quantum dynamics. arXiv:2301.04677 (2023). 63. Lee, J. G., Adelberger, E. G., Cook, T. S., Fleischer, S. M. & Heckel,
39. Oppenheim, J. & Weller-Davies, Z. Covariant path integrals B. R. New test of the gravitational 1/r2 law at separations down to
for quantum fields back-reacting on classical space–time. 52 μm. Phys. Rev. Lett. 124, 101101 (2020).
arXiv:2302.07283 (2023). 64. Kafri, D. & Taylor, J. M. A noise inequality for classical forces.
40. Bohr, N. & Rosenfeld, L. On the question of the measurability of arXiv:1311.4558 [quant-ph] (2013).
electromagnetic field quantities. In Quantum Theory and Mea- 65. Kafri, D., Milburn, G. & Taylor, J. Bounds on quantum commu-
surement (eds Wheeler, J. A. & Wojciech H. Zurek, W. H.). Trans- nication via Newtonian gravity. N. J. Phys. 17, 015006 (2015).
lated by Aage Petersen. Originally published as “Zur Frage der 66. Bose, S. et al. Spin entanglement witness for quantum gravity.
Messbarkeit der Elektromagnetischen Feldgr6ssen" 479–522 Phys. Rev. Lett. 119, 240401 (2017).
(Mat.-fys. Medd Dan. Vid Selsk. 12). (Princeton University Press, 67. Marletto, C. & Vedral, V. Gravitationally induced entanglement
Princeton, [1933] between two massive particles is sufficient evidence of quantum
1983). effects in gravity. Phys. Rev. Lett. 119, 240402 (2017).
41. DeWitt, B. S. Definition of commutators via the uncertainty 68. Marshman, R. J., Mazumdar, A. & Bose, S. Locality and entangle-
principle. J. Math. Phys. 3, 619 (1962). ment in table-top testing of the quantum nature of linearized
42. Gisin, N. Stochastic quantum dynamics and relativity. Helv. Phys. gravity. Phys. Rev. A 101, https://doi.org/10.1103/physreva.101.
Acta 62, 363 (1989). 052110 (2020).
43. Caro, J. & Salcedo, L. Impediments to mixing classical and quan- 69. Pedernales, J. S., Streltsov, K., & Plenio, M. B. (2022). Enhancing
tum dynamics. Phys. Rev. A 60, 842 (1999). Gravitational Interaction between Quantum Systems by a Massive
44. Salcedo, L. Absence of classical and quantum mixing. Phys. Rev. A Mediator. Phys. Rev. Lett. 128, 110401 (2022).
54, 3657 (1996). 70. Carney, D., Müller, H., & Taylor, J. M. Using an Atom Interferometer
45. Sahoo, D. Mixing quantum and classical mechanics and unique- to Infer Gravitational Entanglement Generation. PRX Quantum 2,
ness of Planck’s constant. J. Phys. A: Math. Gen. 37, 997 (2004). 030330 (2021).
46. Terno, D. R. Inconsistency of quantumclassical dynamics, and 71. Kent, A. & Pitala-Garca, D. Testing the nonclassicality of space-
what it implies. Found. Phys. 36, 102 (2006). time: What can we learn from Bellbose et al.-Marletto-Vedral

Nature Communications | (2023)14:7910 22


Article https://doi.org/10.1038/s41467-023-43348-2

experiments? Phys. Rev. D 104, https://doi.org/10.1103/physrevd. 97. Ménoret, V. et al. Gravity measurements below 10-9g
104.126030 (2021). with a transportable absolute quantum gravimeter. Sci. Rep. 8,
72. Christodoulou, M. et al. Locally mediated entanglement through 1 (2018).
gravity from first principles. Phys. Rev. Lett. 130, 100202 (2023). 98. Banks, T., Peskin, M. E. & Susskind, L. Difficulties for the evolution
73. Danielson, D. L., Satishchandran, G. & Wald, R. M. gravitationally of pure states into mixed states. Nucl. Phys. B 244, 125 (1984).
mediated entanglement: Newtonian field vs. gravitons. Phys. Rev. 99. Ghirardi, G. C., Rimini, A. & Weber, T. Unified dynamics for
D 105, 086001 (2022). microscopic and macroscopic systems. Phys. Rev. D 34,
74. Lami, L., Pedernales, J. S. & Plenio, M. B. Testing the quantumness 470 (1986).
of gravity without entanglement. arXiv:2302.03075 (2023). 100. Ballentine, L. Failure of some theories of state reduction. Phys. Rev.
75. Kraus, K. Complementary observables and uncertainty relations. A 43, 9 (1991).
Phys. Rev. D 35, 3070 (1987). 101. Pearle, P., Ring, J., Collar, J. I. & Avignone, F. T. The CSL collapse
76. Kramers, H. A. Brownian motion in a field of force and the diffusion model and spontaneous radiation: an update. Found. Phys. 29,
model of chemical reactions. Physica 7, 284 (1940). 465 (1999).
77. Moyal, J. Stochastic processes and statistical physics. J. R. Stat. 102. Bassi, A., Ippoliti, E. & Vacchini, B. On the energy increase in space-
Soc. Ser. B (Methodological) 11, 150 (1949). collapse models. J. Phys. A: Math. Gen. 38, 8017 (2005).
78. Layton, I., Oppenheim, J. & Weller-Davies, Z. A healthier semi- 103. Adler, S. L. Lower and upper bounds on CSL parameters from
classical dynamics. arXiv:2208.11722 [quant-ph] (2022). latent image formation and IgM heating. J. Phys. A: Math. Theor.
79. Feynman, R. & Vernon, F. The theory of a general quantum system 40, 2935 (2007).
interacting with a linear dissipative system. Ann. Phys. 281, 104. Lochan, K., Das, S. & Bassi, A. Constraining continuous sponta-
547 (2000). neous localization strength parameter λ from standard cosmology
80. Breuer, H.-P. & Petruccione, F. et al. The Theory of Open Quantum and spectral distortions of cosmic microwave background radia-
Systems (Oxford University Press on Demand, 2002). tion. Phys. Rev. D 86, 065016 (2012).
81. Layton, I., Oppenheim, J., Russo, A. & Weller-Davies, Z. The weak 105. Nimmrichter, S., Hornberger, K. & Hammerer, K. Optomechanical
field limit of quantum matter back-reacting on classical space- sensing of spontaneous wave-function collapse. Phys. Rev. Lett.
time. J. High Energy Phys. 2023, 1 (2023). 113, 020405 (2014).
82. Conus, D. & Dalang, R. The non-linear stochastic wave equation in 106. Bahrami, M., Bassi, A. & Ulbricht, H. Testing the quantum super-
high dimensions. Electron. J. Probab. 13, 629 (2008). position principle in the frequency domain. Phys. Rev. A 89,
83. Bassi, A., Lochan, K., Satin, S., Singh, T. P. & Ulbricht, H. Models of 032127 (2014).
wave-function collapse, underlying theories, and experimental 107. Laloë, F., Mullin, W. J. & Pearle, P. Heating of trapped ultracold
tests. Rev. Mod. Phys. 85, 471 (2013). atoms by collapse dynamics. Phys. Rev. A 90, 052119 (2014).
84. Chevalier, H., Paige, A. J. & Kim, M. S. Witnessing the nonclassical 108. Bahrami, M., Paternostro, M., Bassi, A. & Ulbricht, H. Proposal for a
nature of gravity in the presence of unknown interactions. Phys. noninterferometric test of collapse models in optomechanical
Rev. A 102, https://doi.org/10.1103/physreva.102.022428 (2020). systems. Phys. Rev. Lett. 112, 210404 (2014).
85. van de Kamp, T. W., Marshman, R. J., Bose, S. & Mazumdar, A. 109. Goldwater, D., Paternostro, M. & Barker, P. Testing wave-function-
Quantum gravity witness via entanglement of masses: Casimir collapse models using parametric heating of a trapped nano-
screening. Phys. Rev. A 102, 062807 (2020). sphere. Phys. Rev. A 94, 010104 (2016).
86. Toroš, M. et al. Relative acceleration noise mitigation for 110. Tilloy, A. & Stace, T. M. Neutron star heating constraints on wave-
nanocrystal matter-wave interferometry: Applications to function collapse models. Phys. Rev. Lett. 123, 080402 (2019).
entangling masses via quantum gravity. Phys. Rev. Research 3, 111. Donadi, S. et al. Underground test of gravity-related wave function
023178 (2021). collapse. Nat. Phys.17, 74–78 (2021).
87. Chou, C. W., Hume, D. B., Rosenband, T. & Wineland, D. J. Optical 112. Unruh, W. G. & Wald, R. M. Evolution laws taking pure states to
clocks and relativity. Science 329, 1630 (2010). mixed states in quantum field theory. Phys. Rev. D 52, 2176 (1995).
88. Asenbaum, P. et al. Phase shift in an atom interferometer due to 113. Ellis, J., Mavromatos, N. & Nanopoulos, D. V. Quantum-
spacetime curvature across its wave function. Phys. Rev. Lett. 118, gravitational diffusion and stochastic fluctuations in the velocity of
183602 (2017). light. Gen. Relativ. Gravit. 32, 127 (2000).
89. Overstreet, C., Asenbaum, P., Curti, J., Kim, M. & Kasevich, M. A. 114. Parikh, M., Wilczek, F. & Zahariade, G. Signatures of the quanti-
Observation of a gravitational Aharonov–Bohm effect. Science zation of gravity at gravitational wave detectors. Phys. Rev. D 104,
375, 226 (2022). 046021 (2021).
90. Abbott, B. P. et al. Prospects for observing and localizing 115. Verlinde, E. P. & Zurek, K. M. Observational signatures of
gravitational-wave transients with advanced ligo, advanced Virgo quantum gravity in interferometers. Phys. Lett. B 822,
and Kagra. Living Rev. Relativ. 23, 1 (2020). 136663 (2021).
91. McCuller, L. Single-photon signal sideband detection for high- 116. Unruh, W. G. False loss of coherence. In Relativistic Quantum
power Michelson interferometers. arXiv preprint Measurement and Decoherence: Lectures of a Workshop Held at
arXiv:2211.04016 (2022). the Istituto Italiano per gli Studi Filosofici Naples, (eds Breuer, H.-P.
92. Galley, T. D., Giacomini, F., & Selby, J. H. A no-go theorem on the & Petruccione, F.) April 9–10, 1999, 125–140 (Springer, 2000).
nature of the gravitational field beyond quantum theory. Quantum 117. Hall, M. J. W., Cresser, J. D., Li, L. & Andersson, E. Canonical form of
6, 779 (2022). master equations and characterization of non-Markovianity. Phys.
93. Mattingly, D. Modern tests of Lorentz invariance. Living Rev. Rev. A 89, https://doi.org/10.1103/physreva.89.042120 (2014).
Relativ. 8, 1 (2005). 118. Breuer, H.-P., Laine, E.-M., Piilo, J. & Vacchini, B. Colloquium: non-
94. Abbott, P. J. & Kubarych, Z. Mass calibration at NIST in the revised Markovian dynamics in open quantum systems. Rev. Mod. Phys.
SI. Metrolologist 21, 1 (2019). 88, https://doi.org/10.1103/revmodphys.88.021002 (2016).
95. Chao, L. et al. The design and development of a tabletop kibble 119. Layton, I. & Oppenheim, J. The classical-quantum limit.
balance at NIST. IEEE Trans. Instrum. Meas. 68, 2176 (2019). arXiv:2310.18271 (2022).
96. Peters, A., Chung, K. Y. & Chu, S. High-precision gravity mea- 120. Siemon, I., Holevo, A. S. & Werner, R. F. Unbounded generators of
surements using atom interferometry. Metrologia 38, 25 (2001). dynamical semigroups. Open Syst. Inf. Dyn. 24, 1740015 (2017).

Nature Communications | (2023)14:7910 23


Article https://doi.org/10.1038/s41467-023-43348-2

121. Schäfer, G. & Jaranowski, P. Hamiltonian formulation of general the Province of Ontario through the Ministry of Economic Development,
relativity and post-Newtonian dynamics of compact binaries. Liv- Job Creation and Trade.
ing Rev. Relativ. 21, 7 (2018).
122. Oppenheim, J. & Reznik, B. Fundamental destruction of informa- Author contributions
tion and conservation laws. arXiv:0902.2361 [hep-th] (2009) (the J.O., C.S., B.S. and Z.W.D. contributed to this work.
manuscript was never submitted to a journal, but an updated
version is available upon request). Competing interests
123. Ghirardi, G., Rimini, A. & Weber, T. A model for a unified quantum The authors declare no competing interests.
description of macroscopic and microscopic systems. In Quan-
tum Probability and Applications (eds Accardi, L. et al.) (Springer, Additional information
Berlin, 1985). Supplementary information The online version contains
124. Pearle, P. M. Combining stochastic dynamical state vector reduc- supplementary material available at
tion with spontaneous localization. Phys. Rev. A 39, 2277 (1989). https://doi.org/10.1038/s41467-023-43348-2.
125. Ghirardi, G. C., Pearle, P. & Rimini, A. Markov processes in Hilbert
space and continuous spontaneous localization of systems of Correspondence and requests for materials should be addressed to
identical particles. Phys. Rev. A 42, 78 (1990). Jonathan Oppenheim.
126. Hall, M. J. & Reginatto, M. Interacting classical and quantum
ensembles. Phys. Rev. A 72, 062109 (2005). Peer review information Nature Communications thanks Michele
127. Ulmer, W. Deconvolution of a linear combination of Gaussian Arzano, Frank Saueressig and the other, anonymous, reviewer(s) for their
kernels by an inhomogeneous Fredholm integral equation of contribution to the peer review of this work. A peer review file is avail-
second kind and applications to image processing. able.
arXiv:1105.3401 [physics.data-an] (2011).
128. Karolyhazy, F. Gravitation and quantum mechanics of macro- Reprints and permissions information is available at
scopic objects. Il Nuovo Cimento A (1965–1970) 42, 390 (1966). http://www.nature.com/reprints
129. Diósi, L. Models for universal reduction of macroscopic quantum
fluctuations. Phys. Rev. A 40, 1165 (1989). Publisher’s note Springer Nature remains neutral with regard to jur-
130. Penrose, R. On gravity’s role in quantum state reduction. Gen. isdictional claims in published maps and institutional affiliations.
Relativ. Gravit. 28, 581 (1996).
131. Oppenheim, J., Russo, A. & Weller-Davies, Z. Diffeomorphism Open Access This article is licensed under a Creative Commons
invariant classical-quantum path integrals for Nordstrom gravity. Attribution 4.0 International License, which permits use, sharing,
To appear (2023). adaptation, distribution and reproduction in any medium or format, as
long as you give appropriate credit to the original author(s) and the
Acknowledgements source, provide a link to the Creative Commons license, and indicate if
We would like thank Sougato Bose, Joan Camps, Matt Headrick, Isaac changes were made. The images or other third party material in this
Layton, Juan Maldacena, Andrea Russo, Andy Svesko and Bill Unruh for article are included in the article’s Creative Commons license, unless
valuable discussions and Lajos Diósi and Antoine Tilloy for their very indicated otherwise in a credit line to the material. If material is not
helpful comments on an earlier draft of this manuscript. J.O. is sup- included in the article’s Creative Commons license and your intended
ported by an EPSRC Established Career Fellowship, and a Royal Society use is not permitted by statutory regulation or exceeds the permitted
Wolfson Merit Award, C.S. and Z.W.D. acknowledges financial support use, you will need to obtain permission directly from the copyright
from EPSRC. This research was supported by the National Science holder. To view a copy of this license, visit http://creativecommons.org/
Foundation under Grant No. NSF PHY11-25915 and by the Simons licenses/by/4.0/.
Foundation It from Qubit Network. Research at Perimeter Institute is
supported in part by the Government of Canada through the Depart- © The Author(s) 2023
ment of Innovation, Science and Economic Development Canada and by

Nature Communications | (2023)14:7910 24

You might also like