Hawking radiation on a falling lattice
Ted Jacobsona,b∗ and David Mattinglya†
arXiv:hep-th/9908099v2 12 Jan 2000
a
Department of Physics, University of Maryland, College Park, MD 20742-4111, USA
b
Institute for Theoretical Physics, UCSB, Santa Barbara, CA 93106
[4,5]). Quantizing the sound field Unruh argued that the
horizon would radiate thermal phonons, in analogy with
the Hawking effect, at a temperature given by h̄/2π times
the gradient of the flow velocity at the horizon. Since
the sound field has an atomic cutoff, and the physics of
the fluid is completely understood in principle, the fluid
model may provide a source of insight into the Hawking
process.
The atomic cutoff produces dispersion of sound waves
leading to subsonic propagation at high frequencies. It
turns out that this dispersion is all that is needed to obviate the need for a trans-Bohrian reservoir. In the linearized theory, an outgoing short wavelength mode can
be dragged in by the fluid flow and redshifted enough so
that its increased group velocity overcomes the flow and
it escapes to infinity. Investigations of a number of linear
field theory models [6–11] and lattice models [12] incorporating such high frequency dispersion have demonstrated
this mechanism of “mode conversion” and have shown
that the Hawking radiation is recovered.2 The dispersion
in these models is not locally Lorentz-invariant since a
frame is picked out in which to specify which frequencies
are “high”. In the fluid model this is the local rest frame
of the fluid. A real black hole also defines a preferred
frame but it does so in a non-local way. Perhaps quantum gravity produces a dispersion effect related to this
non-local notion of a preferred frame, or perhaps microphysics is just not locally Lorentz invariant. In any case,
it seems worthwhile to understand the mechanism of the
outgoing modes and Hawking radiation in these models
for the hints it may provide about a correct quantum
gravity account.
In [12] a falling lattice model was introduced to address the “stationarity puzzle” [17,8]: how can a low frequency outgoing mode arise from a high frequency ingoing mode when it propagates only in a stationary background spacetime and hence has a conserved Killing frequency? In the continuum based dispersive models there
is no satisfactory resolution of this puzzle. As the outgoing modes are traced backwards in time their incoming
progenitors are squeezed into the short distance regime
Scalar field theory on a lattice falling freely into a 1+1 dimensional black hole is studied using both WKB and numerical approaches. The outgoing modes are shown to arise from
incoming modes by a process analogous to a Bloch oscillation,
with an admixture of negative frequency modes corresponding
to the Hawking radiation. Numerical calculations show that
the Hawking effect is reproduced to within 0.5% on a lattice
whose proper spacing where the wavepacket turns around at
the horizon is ∼ 0.08 in units where the surface gravity is 1.
I. INTRODUCTION
If nothing can emerge from a black hole then where
does the Hawking radiation come from? Relativistic field
theory tells us that it comes from a trans-Planckian reservoir of outgoing modes at the horizon [1]. The existence
of such a trans-Planckian reservoir is doubtful on general
physical grounds. It is the cause of the divergences in
quantum field theory which are expected to be removed
in a quantum theory of gravity, and in particular it would
seem to imply an infinite black hole entropy unless canceled by an infinite negative “bare entropy”.
String theory has produced an account of the Hawking radiation and black hole entropy which requires no
trans-Planckian reservoir, however in that picture instead
of a black hole one has a stringy object in flat spacetime
that arises from a black hole as the string coupling constant tends to zero [2]. Thus gravitational redshift plays
no role in the string calculations and nothing is learned
about the origin of the outgoing modes in a black hole
spacetime.1 If we could follow the theory as the string
coupling becomes strong we would presumably learn how
it is that string theory produces the outgoing black hole
modes.
Absent a quantum gravity description of the Hawking
process, Unruh invented a fluid analog of a black hole [3]
in which an inhomogeneous flow exceeds the long wavelength speed of sound creating a sonic horizon (see also
2
Dispersion can also produce superluminal propagation,
which allows outgoing modes to emerge from behind the horizon. The superluminal case does not seem very healthy from a
fundamental point of view, though it is surely relevant in some
condensed matter analogues of black hole horizons [13–15] and
is capable of producing the usual Hawking spectrum [16].
∗
[email protected]
†
[email protected]
1
The fact that redshifting plays no role in producing the
stringy Hawking radiation is presumably tied to the fact that
the string calculations apply in the “near-extremal” limit that
the surface gravity tends to zero.
1
horizon.
The precise choice of worldlines for the lattice points
should not be important to the leading order Hawking
effect as long as the ground state evolves adiabatically in
the lattice theory. Our numerical results are consistent
with this expectation, in that the particular worldlines
depend on the lattice spacing yet the results converge to
the continuum as the lattice spacing is decreased. Moreover it was shown recently [10], in a continuum model
with high frequency dispersion, that when the preferred
frame is changed from the free-fall frame of the black
hole to that of a conformally related metric there is no
leading order change in the Hawking radiation unless the
acceleration of the preferred frame is very large. Presumably all that matters is that the preferred world lines
flow smoothly across the horizon. From a fundamental
point of view, if there really is a cutoff in some preferred
frame, it is plausible that this frame would coincide with
the cosmic rest frame and would fall across the event
horizons of any black holes that form by collapse.
The remainder of this paper is organized as follows. In
Section II the falling lattice model is set up, in Section
III some results of the WKB method are shown, illustrating the frequency shift and the mode conversion at the
horizon, and the role of the negative frequency branch is
explained. In Section IV the results of the full numerical evolution are presented, and Section V contains a
discussion or the results and directions for further work.
The Appendix describes the finite differencing of the time
derivates used in the numerical calculations.
We use units with h̄ = c = κ = 1, where κ is the
surface gravity.
of the model which is not physically sensible.
A lattice provides a simple model for imposing a physically sensible short distance cutoff. It might seem at
first that the most natural choice would be to preserve
the time translation symmetry of the spacetime with a
static lattice whose points follow accelerated worldlines.
On a static lattice, however, the Killing frequency is conserved, so outgoing modes arise from ingoing modes with
the same frequency, and there is no Hawking radiation.
The in-vacuum therefore evolves to a singular state at the
horizon. If the lattice points are instead freely falling, a
discrete remnant of time translation symmetry can still
be preserved. On such a lattice there is still a frequency
conservation law, and the unsatisfactory resolution of the
stationarity puzzle is that the lattice spacing in this case
goes to zero at infinity (if the lattice points are asymptotically at rest) [12].
If we insist that the lattice spacing asymptotically approaches a fixed constant and that the lattice points are
at rest at infinity, we find a fascinating resolution of the
stationarity puzzle [12]: the lattice cannot have even a
discrete time translation symmetry, since there is a gradual spreading of the lattice points as they fall toward the
horizon. The timescale of this spreading is of order 1/κ
where κ is the surface gravity. This time dependence of
the lattice is invisible to long wavelength modes which
sense only the stationary background metric of the black
hole, but it is quite apparent to modes with wavelengths
of order the lattice spacing. On such a lattice the long
wavelength outgoing modes come from short wavelength
ingoing modes which start out with a high (lattice scale)
Killing frequency which is exponentially redshifted as the
wavepacket propagates to the horizon and turns around.
This behavior of wavepackets in the (time-dependent)
falling lattice model was determined in [12] with the help
of the WKB approximation in which the frequency is
used as a Hamiltonian to solve for the wavepacket trajectory. It was also argued that, since the time dependence
of the lattice is adiabatic for those modes with wavelength short enough to know they are on a lattice, these
modes will remain unexcited on their way to the horizon
and the ground state condition for the Hawking effect
will be met. However, since the wavelength is of order
the lattice spacing and the wavepacket can vary wildly
on the lattice near the horizon, it is not obvious that the
WKB approximation is reliable nor that the adiabatic
argument really applies.
The primary motivation for the work reported here
was to check the WKB and adiabatic assumptions by
carrying out the exact calculation numerically using the
lattice wave equation. We found that indeed the field behaves this way, and the Hawking radiation is recovered
to within half a percent for a lattice spacing δ = 0.002/κ
which corresponds to a proper spacing ∼ 0.08/κ where
the wavepacket turns around at the horizon. We also
studied the deviations from the Hawking effect that arise
as κδ is increased, and our simulations reveal an interesting picture of how the wavepackets turn around at the
II. FALLING LATTICE MODEL
In this section we set up the field theory on a lattice
falling into a black hole in two spacetime dimensions.
A. Black hole spacetime
We assume the spacetime metric is static, so [18] coordinates can be chosen (at least locally) such that the
line element takes the form
2
ds2 = dt2 − dx − v(x)dt .
(2.1)
In these coordinates the Killing vector is given by
χ = ∂t ,
(2.2)
χ2 = 1 − v 2 (x).
(2.3)
with squared norm
For v(x) we choose
v(x) = −
2
vmax
,
cosh(βx)
(2.4)
where vmax and β are positive constants. As |x| →
+∞, v(x) vanishes, so the line element becomes that of
Minkowski spacetime. Provided vmax > 1 there is an ergoregion centered on x = 0 in which the Killing vector
is spacelike, and the boundaries of this region are black
and white hole horizons at
xH = ±β −1 cosh−1 (vmax ).
∂t a/a =
(2.13)
whose maximum outside the √
horizon (cosh(βx)
p > vmax )
2
2
is κ (at the√horizon) if vmax ≥ 2 and κvmax
/2 vmax
−1
<
if vmax ≤ 2. Thus |∂t a/a| ∼ O(κ) everywhere outside
the horizon as long as vmax > 1 is not too close to unity.
(2.5)
The surface gravity of the horizons is given by
p
−2
.
κ = v ′ (xH ) = β 1 − vmax
βvmax sinh(βx)
cosh2 (βx)
B. Lattice
(2.6)
The lattice is defined by discretizing the z coordinate
with spacing ∆z = δ. That is, we have only the discrete
values zm = mδ. We will consider values of κδ which
range from 0.002 to 0.128. To give a feeling for the lattice
we plot in Fig. 1 the (t, x) coordinates of lattice point
worldlines at intervals of 50 lattice points.
The surface gravity sets the length scale for our spacetime, and we will keep it fixed. Hence it will be useful to
employ units for which κ = 1.
We now introduce a freely falling “Gaussian” coordinate z which will be discretized to define a falling lattice.
The worldlines with dx = v(x)dt are geodesics of the
metric (2.1) with proper time t, at rest as |x| → ∞,
and they are orthogonal to the surfaces
R of constant
t. On these curves the quantity t − dx/v is constant, so if z is constant
on these curves we must have
R
W (z(x, t)) = t − dx/v for some function W . We choose
this function
R z so that z = x at t = 0, which implies that
W (z) = − dx/v = sinh(βz)/βvmax , hence
sinh(βz) = sinh(βx) + βvmax t.
30
20
(2.7)
In terms of z, the line element (2.1) takes the form
ds2 = dt2 − a2 (z, t) dz 2 ,
(2.8)
10
(2.9)
0
with
cosh(βz)
.
a(z, t) = p
1 + (sinh(βz) − βvmax t)2
We call a(z, t) the “scale factor”, even though it depends
on z as well as t. It can also be expressed as
v(x(z, t))
a(z, t) =
.
v(z)
0
v(x(z, t))
∂z .
a(z, t)
2
3
4
5
6
FIG. 1. Every 50th lattice
point, t vs. x, for the case
√
δ = 0.004 and vmax = 2. In the black region of the plot
the points are too close to resolve. The horizon is at x ≃ 0.6.
(2.10)
In (t, z) coordinates the Killing vector (2.2) is given by
χ = ∂t − v(z)∂z = ∂t −
1
Since dx = a(z, t)dz at constant t, the separation ∆x
of the lattice points in Fig. 1 gives a direct representation
of the scale factor. The proper lattice spacing is given by
a(z, t)δ which, according to (2.12), grows approximately
linearly with time at the horizon.
Note that there is a region where the lattice spacing
becomes very small. This happens because we chose the
spacing to be uniform at t = 0. If the points did not
bunch up outside the black hole before t = 0, they could
not wind up equally spaced at t = 0, since they are all
on unit energy free-fall trajectories. (The line element is
symmetric under (t, x) → (−t, −x), so the points bunch
up also outside the white hole horizon after t = 0.)
To get a better idea of how sparse the lattice is near
the horizon at the turning point of a typical wavepacket,
(2.11)
The scale factor is unity at t = 0, and outside the black
hole it grows towards the future. It is important to know
how rapidly the scale factor is changing in time, since in
the lattice theory that time dependence is not merely a
coordinate effect and can therefore excite the quantum
field. At the horizon the scale factor takes the value
q
−2
a(zH , t) = 1 + 2κt + κ2 t2 /(1 − vmax
),
(2.12)
so a(zH , t) ∼ κt for κt >
∼ 1. Thus the fractional rate of
change ∂t a/a is of order κ. In general, we have
3
we plot in Fig. 2 the velocity function v (2.4) for every
lattice point near
√ the horizon at time t = 27, for the
case vmax = 2 and κδ = 0.004. In this case there
are only 6 points in the region between v(x) = 0.5 and
the√horizon, and the proper spacing at the horizon is
∼ 2κδt ∼ 0.15/κ.
-1
-2
1
2
hence there is no wave scattering. The discretization
breaks this conformal invariance, however there is still
essentially no scattering on the lattice providing the lattice spacing is much smaller than the radius of curvature.
The reason is that modes with wavelength long compared
to the lattice spacing cannot tell they are not in the continuum, while modes with wavelength comparable to the
lattice spacing cannot tell that the spacetime is curved.
Thus, when we compute the Hawking occupation numbers, there is no “greybody factor”.
3
-0.2
-0.4
-0.6
D. Quantization and the Hawking effect
-0.8
The field is quantized as a collection of self-adjoint operators ϕ̂m (t) satisfying the equation of motion (2.16)
and canonical commutation relations. For each complex
solution to the equations of motion we define the operator a(f ) = (f, ϕ̂) using the inner product (2.17). The
commutation relations are equivalent to the relations
-1
-1.2
-1.4
FIG. √
2. v(x) for every lattice point at t = 27, for the case
vmax = 2 and δ = 0.004.
[a(f ), a† (g)] = (f, g)
(2.18)
for all f and g.
For a positive norm solution p, a(p) acts like a lowering
operator, while for a negative norm solution n, the conjugate n∗ has positive norm, so a(n) = −a† (n∗ ) acts like
a raising operator. The Hilbert space of ingoing modes
is just the Fock space built from solutions with positive
frequency with respect to t (hence positive norm), and
similarly for the outgoing Hilbert space. We assume the
boundary condition that the ingoing positive frequency
field modes are all in their ground state,
C. Scalar field
We consider a massless, minimally coupled scalar field
on the lattice. In the continuum the action would be
Z
√
1
Scont =
d2 x −gg µν ∂µ ϕ∂ν ϕ
2
Z
1
(∂z ϕ)2 . (2.14)
= dt dz (a(z, t)(∂t ϕ)2 −
a(z, t)
where we substituted the metric (2.8) in the second line.
On the lattice we adopt the same action except that the
partial derivative ∂z ϕ is replaced by the finite difference
Dϕm := (ϕm+1 − ϕm )/δ and am (t) is replaced by its
average over the relevant lattice sites:
Z
X
1
2(Dϕm (t))2
2
S=
dt
am (t)(∂t ϕm (t)) −
2
am+1 (t) + am (t)
m
a(pin )|0i = 0.
(2.19)
The occupation number h0|a† (p)a(p)|0i of an outgoing normalized positive frequency wavepacket p in the
in-vacuum |0i is is just minus the norm of the negative
frequency part of the ingoing wavepacket that gives rise
to p. To see this, suppose the ingoing wavepacket q+ +q−
evolves to p according to the field equation (2.16), where
q+ has positive frequency and q− has negative frequency.
∗
Then a(p) = a(q+ ) + a(q− ) = a(q+ ) − a† (q−
), hence
(2.15)
Varying the action (2.15) gives the equation of motion
for ϕm (t),
2Dϕm−1 (t)
= 0. (2.16)
∂t (am (t)∂t ϕm (t)) − D
am+1 (t) + am (t)
h0|a† (p)a(p)|0i = −(q− , q− ),
(2.20)
where we used the in-vacuum condition (2.19) and the
commutation relations (2.18). If p is not normalized then
to obtain the occupation number we must divide by (p, p).
Hawking’s resultR[19] in standard field theory yields, for a
wavepacket p = dω cω exp(−iω(k)t + ikz), the thermal
occupation number
Z
ω|cω |2
NHawking = dω ω/T
,
(2.21)
e H −1
There is a conserved (not positive definite) inner product
between complex solutions:
X
∗
∗
(ψ, ϕ) = i
am (t)(ψm
(t)∂t ϕm (t) − ϕm (t)∂t ψm
(t)).
m
(2.17)
In two dimensions the continuum action (2.14) is conformally invariant, and all metrics are conformally flat,
where TH = κ/2π is the Hawking temperature.
4
(in the free fall frame) changes sign and the wavevector is
equivalent to +π/δ. From there it continues to decrease,
winding up small and positive. This reversal of group
velocity by a monotonic change of the wavevector on a
lattice is analogous to the Bloch oscillation of an electron
in a crystal acted on by a uniform electric field.
Thus the interval (−π/δ, −|kc |) is mapped onto the interval (0, π/δ). The map is onto since every outgoing
wavevector arises from some ingoing wavevector. This
dynamical stretching of the k-space interval of length
π/δ − |kc | to one of length π/δ is not problematic in a
particle phase flow, but if we think about the wavepackets that are being approximated by this WKB analysis it
is disturbing since there really are more outgoing modes
than ingoing modes which produced them. It seems the
wave evolution is somehow not conserving the number of
wave degrees of freedom.
The resolution of this puzzle is the essence of the Hawking effect. The wave evolution mixes positive and negative frequency modes, so a purely positive frequency outgoing wavepacket arises from a mixture of positive and
negative frequency ingoing wavepackets. WKB evolution breaks down at the turing point, hence it misses this
mode mixing.
An example of a positive frequency WKB trajectory is
shown in Fig. 4. The graph shows several different quantities as functions of t (in units with κ = 1): the static
position coordinate of the wavepacket x and the horizon
xH (solid lines), the free-fall frequency ω (3.1) (dashed
line), the Killing frequency ωK (3.2) (dot-dashed line),
and 50 sin(kδ/2) (sparsely-dashed line). The positions
are scaled up by a factor of 10 so they can be seen on
the same graph. In
√ this example the lattice spacing is
δ = 0.004, vmax = 2, and the trajectory has z(34) = 6
and k(34) = 1.
III. WKB ANALYSIS
The motion of wavepackets in the falling lattice model
can be studied using the WKB approximation [12]. This
amounts to using Hamilton’s equations for the position
z and momentum k with a Hamiltonian determined by
the dispersion relation
H(z, k, t) = ω = ±
2/δ
sin(kδ/2)
a(z, t)
(3.1)
which is obtained by inserting a mode of the form
exp(−iωt + ikz) into the field equation and keeping only
the terms with the highest derivatives or finite differences
of the field [20]. A plot of this dispersion relation is shown
in Fig. 3 for a(z, t) = 1. Wavevectors differing by 2π/δ
are equivalent, so only the range (−π/δ, π/δ) (the “Brillouin zone”) is plotted.
2
1
-3
-2
-1
1
2
3
-1
-2
FIG. 3. Dispersion relation ωδ = ±2 sin(kδ/2) plotted vs.
kδ.
The frequency ω is not conserved, even when kδ ≪ 1,
simply because it is not the Killing frequency. Using
(2.11) we see that in the continuum the Killing frequency
would be given by
50
40
ωK
v(x(z, t))
=ω+
k,
a(z, t)
(3.2)
sin(kδ/2)
30
with which the dispersion relation can be expressed as
a(z, t)ωK − v(x)k = (2/δ) sin(kδ/2).
20
free-fall frequency
Killing frequency
packet trajectory
(3.3)
10
Note however that this relation is only valid when kδ
is not too large, otherwise the fact that the Killing frequency involves a partial derivative in a direction that is
not along a lattice point worldline renders the relation
meaningless.
It was shown in [12] that the WKB trajectory of an ingoing positive frequency wavepacket bounces off the horizon and comes back out provided the ingoing wavevector is greater in magnitude than some critical value |kc |.
(This critical value depends on the time and place from
which the packet is launched.) The wavevector evolves
from negative values to positive values by decreasing until it is less than −π/δ, at which point the group velocity
horizon
20
22.5
25
27.5
30
32.5
Time
FIG. 4. WKB evolution.
Notice first the behavior of sin(kδ/2). Backwards in
time this starts out very small, rises to the maximum,
and comes back down but not to zero. The group velocity in the frame falling with the lattice vanishes at the
top, when kδ/2 = π/2, corresponding to a wavelength
equal to two lattice units. This rise and fall happens
via a monotonic change of k, and is the Bloch oscillation
5
mentioned above.
The wavepacket that ends up with ω = ωK = 1 at
late times starts out at early times with ω = ωK ∼ 43.
The maximum possible lattice frequency is 2/δ = 500
in this case, so the ingoing frequency is certainly “high”.
Whereas the free-fall frequency continues to redshift (forward in time) as the wavepacket climbs away from the
horizon, the redshifting of the Killing frequency is essentially complete by the time the wavepacket reaches the
turning point. Thus, past the turning point, each Killing
frequency component propagates autonomously, more or
less as it would in the continuum. As described in the
Introduction, this, together with the fact that the evolution of ω at the lattice scale is adiabatic, is the reason
why the usual Hawking effect is reproduced.
B. Wavepackets and parameters
If the Killing frequency were conserved as it is in
the dispersive continuum models the entire Hawking
spectrum could be probed just by propagating one
wavepacket and decomposing it into is Fourier components, each of which would pick up a negative frequency
part appropriate for the corresponding frequency. This
is how Unruh checked the spectrum in his model [6].
Since the Killing frequency is not conserved on our lattice
we cannot avail ourselves of this option.3 Thus instead
we compute the occupation numbers for a sequence of
wavepackets ϕ
es (k) with different profiles:
ϕ
es (k) = k exp[−(k − 0.05s)2 ],
s = 1, . . . , 9
(4.1)
(in units with κ = 1, as usual). The form of these
wavepackets was dictated by the need to have them well
enough localized to be contained in the flat region of
the spacetime (so we could construct the corresponding positive frequency initial data) while at the same
time containing enough power at low frequencies of order κ to have a measurable Hawking occupation number. To shift the wavepacket to the desired starting position the Fourier transform (4.1) was multiplied by a
phase factor exp(−ikzinitial). The positive frequency initial data was constructed by evolving each Fourier component one time step ∆t by multiplying with the phase
factor exp(−iω(k)∆t) and Fourier transforming the result back into position space.
The time dependence of the lattice (illustrated in Fig.
1) means that the results will not be independent of
the “launch time” of the wavepacket from a given location. If we start our backwards evolution too far in
the future then when the wavepacket nears the horizon
the lattice points will be pathologically sparse. On the
other hand if we launch backwards at too early a time
then the wavepacket can become partially trapped in the
region where the lattice spacing gets very small. This
happens because a sufficiently high frequency wave is
turned around when it tries to propagate into a region
of larger lattice spacing. We encountered this “channeling” phenomenon in our simulations, and avoided it
just by launching later. The runs reported here were all
done with wavepackets that reached the horizon around
t = 27, and were launched from far enough away to be
contained in the flat region of the spacetime. Typically,
IV. LATTICE WAVE EQUATION ANALYSIS
In this section we discuss the results of our numerical
calculations. The basic method is to take an outgoing
positive frequency wavepacket and evolve it backwards
in time. After it has “bounced” from the horizon and
propagated back into the flat region of the spacetime we
decompose it into positive and negative frequency parts.
Since it is a purely left moving wavepacket at this stage
the negative frequency part is just the part composed
of positive wavevectors. The occupation number (2.20)
of the original outgoing packet is then given by minus
the norm of this negative frequency part divided by the
total norm. We typically computed the total norm in
position space using (2.17). Since the negative frequency
part ϕ(−) was projected out in the wavevector space we
usedPan alternate expression for its norm, (ϕ(−) , ϕ(−) ) =
−2 k>0 |ω(k)||ϕ
e(−) (k)|2 , where the frequency ω(k) is
given by (3.1).
A. Numerical issues
Our system of equations (2.16) is a coupled set of ordinary differential equations, which is to be solved in the
limit that time is continuous but the spatial lattice is
fixed. In order to satisfy the Courant stability criterion
the time step ∆t must satisfy ∆t < a(z, t)δ. (In this
1+1 dimensional setting the Courant condition coincides
with the condition that causality be respected by the differencing scheme.) In the asymptotic region a(z, t) = 1
so the stability condition there is ∆t < δ. For a fixed
∆t/δ the general condition would be violated wherever
a(z, t) < ∆t/δ. Since this only happens deep behind the
horizon (for t > 0) we just modified a(z, t) inside the
black hole so that it is never less than ∆t/δ. We used a
time discretization scheme (written out in the Appendix)
with errors of order (∆t)2 , and found that ∆t = 0.4δ was
adequately small to obtain very good accuracy.
3
We studied the possibility that the WKB evolution could
be used to establish a mapping between in and out frequencies, thus allowing us to check the entire spectrum of particle
creation just by propagating a single wavepacket. This turns
out to be a flawed idea however since the modes are mixed
by the evolution and, moreover, such a map misses the role
of the negative frequency piece.
6
the launch time was around t = 56 when the wavepacket
was centered at x = 28.
All of the calculations reported here were done using
√
the line element (2.1), with vmax (2.4) ranging from 2
to 30.
A sample of the comparisons for κδ = 0.002 is listed
in Table I. The typical relative difference in this case is
of order half a percent.
3. Lattice dependence
C. Results
To study the dependence on the lattice we ran the
s = 1 wavepacket backward at several different values of
the lattice
spacing and for several values of vmax (2.4)
√
from 2 to 30. The change of vmax produces only a
rather small change in the shape of the function v(x)
outside the horizon, however it produces a shift of the
position of the horizon (2.5). For large vmax the horizon
is given approximately by xH ≃ ln(2vmax ).√The horizon
position ranges from about 0.6 for vmax = 2 to about 4
for vmax = 30. Thus a wavepacket launched backward in
time from a given position at a given time will reach the
horizon later in time for larger vmax , so that the proper
lattice spacing at the horizon will be larger, though not
by a very large amount for the wavepackets considered
here.
In Fig. 5 we plot the occupation numbers for lattice
spacings δ ranging from 0.002√to 0.128, and for several
values of vmax (2.4), vmax = 2, 2, 2.5, 3, 4, 5, 10, 20, 30.
In all cases the late time wavepacket was launched from
the same position and time. All wavepackets had the
same envelope for their Fourier transform, however for
different lattice spacings the wavepacket is necessarily
different.
1. Generalities
The behavior of a typical wavepacket throughout the
process of bouncing off the horizon is illustrated in Fig.
6. The real part of the wavepacket is plotted vs. the
static coordinate x at several different times. Following
backwards in time, the wave starts to squeeze up against
the horizon and then a trailing dip freezes and develops
oscillations that grow until they balloon out, forming into
a compact high frequency wavepacket that propagates
neatly away from the horizon backwards in time.
This ingoing wavepacket contains both positive and
negative frequency components, in just the right combination to produce only an outgoing wave when sent in
towards the black hole, with no wave propagating across
the horizon. To illustrate this we have taken one of these
ingoing wavepackets, decomposed it into its positive and
negative frequency parts, and sent them separately back
toward the horizon forward in time (Fig. 7). The negative frequency part (which is the smaller of the two since
the surface gravity is fairly low compared to the typical
frequencies in the outgoing wavepacket), mostly crosses
the horizon, with just a bit bouncing back out. The positive frequency part mostly bounces, with just a little bit
going into the black hole.
Occupation #
0.025
2. Spectrum
The “observed” occupation numbers for the wavepackets (4.1) all agree well with the integrated thermal predictions (2.21) of the Hawking effect provided the lattice
spacing is not too large.
Thermal
0.01563
0.01409
0.01266
0.01135
0.01014
0.00903
0.00801
0.00709
0.00625
Lattice
0.01557
0.01404
0.01262
0.01130
0.01009
0.00899
0.00797
0.00705
0.00622
Thermal Value
0.015
0.01
0.005
0
0
0.05
0.1
0.15
Delta
TABLE I. Comparison of the black hole radiation on the
lattice to the thermal Hawking occupation√numbers (2.21) for
the wavepackets (4.1) for the case vmax = 2 and κδ = 0.002.
s
1
2
3
4
5
6
7
8
9
0.02
FIG. 5. Occupation numbers for s = 1 wavepacket for various values of δ and vmax . Higher occupation numbers correspond to larger values of vmax .
Rel. Diff.
0.0038
0.0035
0.0034
0.0041
0.0045
0.0039
0.0051
0.0052
0.0046
For all values of vmax , the occupation number converges to the Hawking prediction as the lattice spacing δ
decreases.
The pattern of deviations from the thermal prediction
appears to depend quite strongly on the value of vmax , a
result which we do not understand at present. The best
we can do is to list various effects which would contribute
to the deviations: (i) the WKB turning point is moving away from the horizon as δ grows, so one might ex-
7
pect that the effective surface gravity v ′ (xt.p. ) and hence
the effective Hawking temperature felt by the wavepacket
would decrease; (ii) at larger vmax the proper lattice spacing at the horizon is larger at the time this particular
wavepacket reaches the horizon; (iii) the lattice points
are rather sparse near the horizon for the larger values
of δ, and their precise positions will be significantly different for different values of vmax . Effect (i) would lead
to a decrease in the occupation number, while one might
expect that effect (ii) would lead to a relative increase
due to the extra particle creation associated with time
dependence of the lattice. Perhaps what we are seeing is
just a delicate balance between these two effects.
in others, which requires at least two dimensions to be
possible. At the atomic level such a volume-preserving
flow involves erratic motions of individual atoms. One
possible improvement of the lattice model is to make a
lattice that mimics this sort of volume preserving flow. It
is not clear whether the motions of the lattice points can
be slow enough to be adiabatic on the timescale of the
high frequency lattice modes. If they cannot, then the
time dependence of this erratic lattice background will
excite the quantum vacuum.
In a fluid the lattice is a part of the system, not just
a fixed background. The average flow could be adiabatic
for the fully coupled ground state of the system but not
for the field theory of the perturbations on the “background lattice”. Similar comments apply in quantum
gravity: surely if in-out mode conversion is at play the
incoming high frequency modes are strongly coupled to
the quantum gravitational vacuum. Ideally, therefore, we
should try to find a model in which the background is not
decoupled from the perturbations.
As a first step in this direction one could study a one
dimensional model in which the lattice points are nonrelativistic point masses, coupled to each other by nearest neighbor interactions, and “falling” or propagating in
a background potential (with or without periodic boundary conditions). The perturbations of such a lattice are
the phonon field, and the back reaction to the Hawking
radiation is included (although the background potential
is fixed). In a model like this one could presumably follow
in detail the nonlinear origin of the outgoing modes and
the transfer of energy from the mean flow to the thermal
radiation.
V. DISCUSSION
We have shown that linear field theory on a falling lattice reproduces the continuum Hawking effect with high
fidelity, thus verifying earlier expections [12]. For the
smallest lattice spacing we studied, δ = 0.002/κ, the deviations from the thermal particle creation in wavepackets composed of frequencies ω <
∼ O(κ) were of order
half a percent. This is remarkable, particularly since the
proper lattice spacing encountered by the wavepacket at
the horizon was ∼ 0.08/κ. When κδ is pushed to higher
values eventually we see significant deviations from the
Hawking prediction.
There are several directions in which this work could
be developed. One idea is to include self-interactions for
the quantum field theory. There are general arguments
[21,22] that the thermal Hawking effect holds also for interacting fields. There are also perturbative calculations
[23] and calculations exploiting exact conformal invariance of certain 1+1 dimensional field theories [24]. Also
the “hadronization” of Hawking radiation from primordial black holes has been analyzed by matching to high
energy collision data [25], but the physics of how the thermal Hawking radiation is “dressed” by interactions as it
redshifts away from the horizon has not been studied. In
a 1+1 dimensional model it should not be prohibitively
difficult to numerically study Hawking radiation of interacting fields, now that we see it can be done on a lattice.
The falling lattice model has provided a satisfactory
mechanism—the Bloch oscillation—for how to get an
outgoing mode from an ingoing mode in a stationary
background, but there is a serious flaw in the picture:
the lattice is constantly expanding. In the fluid model by
contrast the lattice of atoms maintains a uniform average
density. In a fundamental theory we might also expect
the scale of graininess of spacetime to remain fixed at,
say, the Planck scale or the string scale (since presumably the graininess would define this scale) rather than
expand. Can the falling lattice model be improved to
share this feature?
A fluid maintains uniform density in an inhomogeneous
flow by compressing in some directions and expanding
ACKNOWLEDGEMENTS
We are grateful to Matt Choptuik and David Garfinkle
for advice on numerical matters and to Steve Corley for
many helpful discussions. This work was supported in
part by the National Science Foundation under grants
No. PHY98-00967 at the University of Maryland and
PHY94-07194 at the Institute for Theoretical Physics.
8
APPENDIX: FINITE DIFFERENCING SCHEME
The equation of motion (2.16) for ϕm (t) is
∂t am (t)∂t ϕm (t) = D
2Dϕm−1 (t)
am+1 (t) + am (t)
.
We used a time discretization scheme for the left hand side arising from the replacement of f˙ by
h
i
f (t + ∆t/2) − f (t − ∆t/2) /∆t = f˙ + (1/24)f (3)(∆t)2 + O((∆t)4 ),
(A1)
(A2)
which yields
h
i
∂t (am (t)∂t ϕm (t)) → a(n + 1/2) ϕ(n + 1) − ϕ(n) − a(n − 1/2) ϕ(n) − ϕ(n − 1) /(∆t)2 ,
(A3)
where the spatial index is omitted and the argument n stands for n∆t. We could have used this replacement, since
a(z, t) isa prescribed function
that can be calculated at any t, but for some reason we substituted a(n ± 1/2) by the
average a(n) + a(n ± 1) /2. This yields the approximation
h
i
a(n) + a(n + 1) ϕ(n + 1) − ϕ(n) − a(n) + a(n − 1) ϕ(n) − ϕ(n − 1) /2(∆t)2
1
1
1
1
aϕ(4) + ȧϕ(3) + äϕ̈ + a(3) ϕ̇ (n)(∆t)2 + O((∆t)4 ).
= ∂t (a∂t ϕ)(n) +
12
6
4
4
(A4)
(A5)
With (A4) in place of the left hand side of (A1) we solve for ϕ(n − 1) in terms of ϕ(n) and ϕ(n + 1) to obtain an
equation for evolving ϕ backwards in time by one time step given the following two steps.
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
T. Jacobson, Phys. Rev. D44, 1731 (1991).
For a review see A.W. Peet, Class. Quant. Grav. 15, 3291 (1998).
W.G. Unruh, Phys. Rev. Lett. 46, 1351 (1981).
V. Moncrief, Astrophys. J, 235, 1038 (1980).
M. Visser, Class. Quantum Grav. 15, 1767 (1998).
W.G. Unruh, Phys. Rev. D 51, 2827 (1995).
R. Brout, S. Massar, R. Parentani and Ph. Spindel, Phys. Rev. D 52, 4559 (1995).
S. Corley and T. Jacobson, Phys. Rev. D 54,1568 (1996).
S. Corley, Phys. Rev. D 57, 6280 (1998).
Y. Himemoto and T. Tanaka, “A generalization of the model of Hawking radiation with modified high frequency dispersion
relation”, gr-qc/9904076.
H. Saida and M. Sakagami, “Black hole radiation with high frequency dispersion”, gr-qc/9905034.
S. Corley and T. Jacobson, Phys. Rev. D 57, 6269 (1998).
T.A. Jacobson and G.E. Volovik, Phys. Rev. D 58, 064021 (1998); JETP Lett. 68, 874 (1998).
N.B. Kopnin and G.E. Volovik, Phys. Rev. B 57, 8526 (1998); Phys. Rev. Lett. 79, 1377 (1997).
G.E. Volovik, Pis’ma Zh. Eksp. Teor. Fiz. 69, 662 (1999) [JETP Lett. 69, 705 (1999).]
S. Corley, Phys. Rev. D 57, 6280 (1998).
T. Jacobson, Phys. Rev. D 53 (1996) 7082.
See e.g. Appendix A of ref. [12].
S.W. Hawking, Nature (London) 248 (1974) 30; Commun. Math. Phys. 43 (1975) 199.
The Hamiltonian technique was first applied to dispersive models of the Hawking effect in ref. [7].
G.W. Gibbons and M.J. Perry, Proc. Roy. Soc. Lond., Series A 358, 467-94 (1977).
W.G. Unruh and N. Weiss, Phys. Rev. D 29, 1656 (1984).
D.A. Leahy and W.G. Unruh, Phys. Rev. D 28, 694 (1983).
N.D. Birrell and P.C.W. Davies, Phys. Rev. D 18, 4408 (1978).
J.H. MacGibbon and B.R. Webber, Phys. Rev. D 41, 3052 (1990); J.H. MacGibbon, Phys. Rev. D 44, 376 (1991); F.
Halzen, E. Zas, J.H. MacGibbon, and T.C. Weekes, Nature 353, 807 (1991).
9
√
FIG. 6. A typical wavepacket evolution. Here vmax = 2, κδ = 0.004, and the wavepacket (4.1) has s = 1. The ocsillations
of the incoming wavepacket are too dense to resolve in the plots.
10
FIG. 7. The positive (top row) and negative (bottom row) frequency parts of the ancestor of an outgoing positive frequency
wavepacket, sent back towards the black hole. Note the vertical scale is different in the two plots. The wavepackets crossing the
horizon in the two cases are negatives of each other since they must cancel to produce only the outgoing wavepacket we started
with. The parts that go through the horizon appear on the right due to periodic boundary conditions. The high frequency
ocsillations are too dense to resolve in the plots.
11