Theorems which give fundamental limits on quantum evolution
Quantum speed limit theorems are quantum mechanics theorems concerning the orthogonalization interval, the minimum time for a quantum system to evolve between two orthogonal states , also known as the quantum speed limit .
Consider an initial pure quantum state expressed as a superposition of its energy eigenstates
|
ψ
(
0
)
⟩
=
∑
n
c
n
|
E
n
⟩
{\displaystyle \left|\psi (0)\right\rangle =\sum _{n}c_{n}\left|E_{n}\right\rangle }
.
If the state
|
ψ
(
0
)
⟩
{\displaystyle \left|\psi (0)\right\rangle }
is let to evolve for an interval
δ
t
{\displaystyle \delta t}
by the Schrödinger equation it becomes
|
ψ
(
δ
t
)
⟩
=
∑
n
c
n
e
−
i
E
n
δ
t
ℏ
|
E
n
⟩
{\displaystyle \left|\psi (\delta t)\right\rangle =\sum _{n}c_{n}e^{-i{\frac {E_{n}\delta t}{\hbar }}}\left|E_{n}\right\rangle }
,
where
ℏ
=
h
2
π
{\displaystyle \hbar ={\frac {h}{2\pi }}}
is the reduced Planck constant .
If the initial state
|
ψ
(
0
)
⟩
{\displaystyle \left|\psi (0)\right\rangle }
is orthogonal to the evolved state
|
ψ
(
δ
t
)
⟩
{\displaystyle \left|\psi (\delta t)\right\rangle }
then
⟨
ψ
(
0
)
|
ψ
(
δ
t
)
⟩
=
0
{\displaystyle \left\langle \psi (0)|\psi (\delta t)\right\rangle =0}
and the minimum interval
δ
t
⊥
{\displaystyle \delta t_{\perp }}
required to achieve this condition is called the orthogonalization interval[ 1] or time.[ 2]
Mandelstam-Tamm theorem
The Mandelstam-Tamm theorem[ 1] states that
δ
E
δ
t
⊥
≥
ℏ
π
2
{\displaystyle \delta E\delta t_{\perp }\geq \hbar {\frac {\pi }{2}}}
,
where
(
δ
E
)
2
=
⟨
ψ
|
H
2
|
ψ
⟩
−
(
⟨
ψ
|
H
|
ψ
⟩
)
2
=
1
2
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
(
E
n
−
E
m
)
2
{\displaystyle (\delta E)^{2}=\left\langle \psi |H^{2}|\psi \right\rangle -(\left\langle \psi |H|\psi \right\rangle )^{2}={\frac {1}{2}}\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}(E_{n}-E_{m})^{2}}
,
is the variance of the system's energy and
H
{\displaystyle H}
is the Hamiltonian operator.
The theorem is named after Leonid Mandelstam and Igor Tamm .
In this case, quantum evolution is independent of the particular Hamiltonian used to transport the quantum system along a given curve in the projective Hilbert space; it is the distance along this curve measured by the Fubini-Study metric .[ 3]
Proof
We want to find the smallest interval
δ
t
⊥
{\displaystyle \delta t_{\perp }}
such that
|
S
(
δ
t
⊥
)
|
2
=
|
⟨
ψ
(
0
)
|
ψ
(
δ
t
⊥
)
⟩
|
2
=
0
{\displaystyle |S(\delta t_{\perp })|^{2}=|\left\langle \psi (0)|\psi (\delta t_{\perp })\right\rangle |^{2}=0}
.
We note[ 2] that
|
S
(
δ
t
)
|
2
=
|
⟨
ψ
(
0
)
|
ψ
(
δ
t
)
⟩
|
2
=
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
e
−
i
δ
t
ℏ
(
E
n
−
E
m
)
=
=
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
cos
(
δ
t
ℏ
(
E
n
−
E
m
)
)
,
{\displaystyle {\begin{aligned}|S(\delta t)|^{2}&=|\left\langle \psi (0)|\psi (\delta t)\right\rangle |^{2}=\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}e^{-i{\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)}=\\&=\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}\cos \left({\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)\right),\end{aligned}}}
using Euler's formula and noting that the sine function is odd. Then
|
S
(
δ
t
)
|
2
≥
1
−
4
π
2
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
δ
t
ℏ
(
E
n
−
E
m
)
sin
(
δ
t
ℏ
(
E
n
−
E
m
)
)
−
2
π
2
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
(
δ
t
ℏ
(
E
n
−
E
m
)
)
2
{\displaystyle {\begin{aligned}|S(\delta t)|^{2}&\geq 1-{\frac {4}{\pi ^{2}}}\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}{\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)\sin \left({\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)\right)\\&-{\frac {2}{\pi ^{2}}}\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}\left({\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)\right)^{2}\end{aligned}}}
,
since
cos
(
x
)
≥
1
−
4
π
2
x
sin
(
x
)
−
2
π
2
x
2
{\displaystyle \cos(x)\geq 1-{\frac {4}{\pi ^{2}}}x\sin(x)-{\frac {2}{\pi ^{2}}}x^{2}}
,
∀
x
∈
R
{\displaystyle \forall x\in \mathbb {R} }
.
We note that
d
|
S
(
δ
t
)
|
2
d
δ
t
=
−
∑
n
,
m
|
c
n
|
2
|
c
m
|
2
sin
(
δ
t
ℏ
(
E
n
−
E
m
)
)
E
n
−
E
m
ℏ
{\displaystyle {\frac {d|S(\delta t)|^{2}}{d\delta t}}=-\sum _{n,m}|c_{n}|^{2}|c_{m}|^{2}\sin \left({\frac {\delta t}{\hbar }}\left(E_{n}-E_{m}\right)\right){\frac {E_{n}-E_{m}}{\hbar }}}
.
Thus
|
S
(
δ
t
)
|
2
≥
1
+
4
π
2
δ
t
d
|
S
(
δ
t
)
|
2
d
δ
t
−
1
π
2
(
2
δ
t
ℏ
δ
E
)
2
{\displaystyle |S(\delta t)|^{2}\geq 1+{\frac {4}{\pi ^{2}}}\delta t{\frac {d|S(\delta t)|^{2}}{d\delta t}}-{\frac {1}{\pi ^{2}}}\left({\frac {2\delta t}{\hbar }}\delta E\right)^{2}}
.
Since
|
S
(
δ
t
)
|
2
≥
0
{\displaystyle |S(\delta t)|^{2}\geq 0}
then
d
|
S
(
δ
t
)
|
2
d
δ
t
=
0
{\displaystyle {\frac {d|S(\delta t)|^{2}}{d\delta t}}=0}
if
S
(
δ
t
)
=
0
{\displaystyle S(\delta t)=0}
. So the second term vanishes for
δ
t
=
δ
t
⊥
{\displaystyle \delta t=\delta t_{\perp }}
and
0
≥
1
−
1
π
2
4
δ
t
⊥
2
ℏ
2
(
δ
E
)
2
{\displaystyle 0\geq 1-{\frac {1}{\pi ^{2}}}{\frac {4\delta t_{\perp }^{2}}{\hbar ^{2}}}\left(\delta E\right)^{2}}
.
For this bound to become an equality we demand
cos
(
x
)
=
1
−
4
π
2
x
sin
(
x
)
−
2
π
2
x
2
{\displaystyle \cos(x)=1-{\frac {4}{\pi ^{2}}}x\sin(x)-{\frac {2}{\pi ^{2}}}x^{2}}
, that is
x
=
0
{\displaystyle x=0}
or
x
=
±
π
{\displaystyle x=\pm \pi }
. Thus
δ
t
⊥
ℏ
(
E
n
−
E
m
)
=
0
or
δ
t
⊥
ℏ
(
E
n
−
E
m
)
=
±
π
∀
n
,
m
,
c
n
≠
0
,
c
m
≠
0
{\displaystyle {\frac {\delta t_{\perp }}{\hbar }}\left(E_{n}-E_{m}\right)=0\quad {\text{or}}\quad {\frac {\delta t_{\perp }}{\hbar }}\left(E_{n}-E_{m}\right)=\pm \pi \quad \forall n,m,c_{n}\neq 0,c_{m}\neq 0}
,
which holds for only two energy eigenstates
E
0
=
0
{\displaystyle E_{0}=0}
and
E
1
=
±
π
ℏ
δ
t
⊥
{\displaystyle E_{1}=\pm {\frac {\pi \hbar }{\delta t_{\perp }}}}
. Thus, the only state that attains this bound is a two-level pure quantum state (qubit ) in an equal superposition
|
ψ
q
⟩
=
1
2
(
e
i
φ
0
|
0
⟩
+
e
i
φ
1
|
±
π
ℏ
δ
t
⊥
⟩
)
{\displaystyle \left|\psi _{q}\right\rangle ={\frac {1}{\sqrt {2}}}\left(e^{i\varphi _{0}}\left|0\right\rangle +e^{i\varphi _{1}}\left|\pm {\frac {\pi \hbar }{\delta t_{\perp }}}\right\rangle \right)}
of energy eigenstates
|
E
0
⟩
{\displaystyle \left|E_{0}\right\rangle }
and
|
E
1
⟩
{\displaystyle \left|E_{1}\right\rangle }
, unique up to degeneracy of the energy level
E
1
{\displaystyle E_{1}}
and arbitrary phase factors
φ
0
{\displaystyle \varphi _{0}}
,
φ
1
{\displaystyle \varphi _{1}}
of the eigenstates.[ 2]
Margolus–Levitin theorem
The Margolus–Levitin theorem[ 4] states that
E
a
v
g
δ
t
⊥
≥
ℏ
π
2
{\displaystyle E_{avg}\delta t_{\perp }\geq \hbar {\frac {\pi }{2}}}
,
where
E
a
v
g
=
⟨
ψ
|
H
|
ψ
⟩
=
∑
n
|
c
n
|
2
E
n
{\displaystyle E_{avg}=\left\langle \psi |H|\psi \right\rangle =\sum _{n}|c_{n}|^{2}E_{n}}
,
is the system's average energy and
H
{\displaystyle H}
is the Hamiltonian operator, such that
H
{\displaystyle H}
does not depend on time;
H
{\displaystyle H}
has zero ground state energy.
The theorem is named after Norman Margolus and Lev B. Levitin .
Proof
Graphs of trigonometric functions used in inequalities of Mandelstam-Tamm and Margolus–Levitin theorems.
We want to find the smallest interval
δ
t
⊥
{\displaystyle \delta t_{\perp }}
such that
S
(
δ
t
⊥
)
=
⟨
ψ
(
0
)
|
ψ
(
δ
t
⊥
)
⟩
=
∑
n
|
c
n
|
2
e
−
i
E
n
δ
t
⊥
ℏ
=
0
{\displaystyle S(\delta t_{\perp })=\left\langle \psi (0)|\psi (\delta t_{\perp })\right\rangle =\sum _{n}|c_{n}|^{2}e^{-i{\frac {E_{n}\delta t_{\perp }}{\hbar }}}=0}
.
We note that[ 2]
Re
(
S
(
δ
t
)
)
=
∑
n
|
c
n
|
2
cos
(
E
n
δ
t
ℏ
)
≥
≥
∑
n
|
c
n
|
2
(
1
−
2
π
E
n
δ
t
ℏ
−
2
π
sin
(
E
n
δ
t
ℏ
)
)
=
=
∑
n
|
c
n
|
2
−
2
δ
t
π
ℏ
∑
n
|
c
n
|
2
E
n
−
2
π
∑
n
|
c
n
|
2
sin
(
E
n
δ
t
ℏ
)
=
=
1
−
2
δ
t
π
ℏ
E
a
v
g
+
2
π
Im
(
S
(
δ
t
)
)
{\displaystyle {\begin{aligned}{\text{Re}}(S(\delta t))&=\sum _{n}|c_{n}|^{2}\cos \left({\frac {E_{n}\delta t}{\hbar }}\right)\geq \\&\geq \sum _{n}|c_{n}|^{2}\left(1-{\frac {2}{\pi }}{\frac {E_{n}\delta t}{\hbar }}-{\frac {2}{\pi }}\sin \left({\frac {E_{n}\delta t}{\hbar }}\right)\right)=\\&=\sum _{n}|c_{n}|^{2}-{\frac {2\delta t}{\pi \hbar }}\sum _{n}|c_{n}|^{2}E_{n}-{\frac {2}{\pi }}\sum _{n}|c_{n}|^{2}\sin \left({\frac {E_{n}\delta t}{\hbar }}\right)=\\&=1-{\frac {2\delta t}{\pi \hbar }}E_{avg}+{\frac {2}{\pi }}{\text{Im}}(S(\delta t))\end{aligned}}}
,
as
cos
(
x
)
≥
1
−
2
π
x
−
2
π
sin
(
x
)
,
∀
x
≥
0
{\displaystyle \cos(x)\geq 1-{\frac {2}{\pi }}x-{\frac {2}{\pi }}\sin(x),\forall x\geq 0}
.
Since
S
(
δ
t
⊥
)
=
0
{\displaystyle S(\delta t_{\perp })=0}
requires
Re
(
S
(
δ
t
⊥
)
)
=
Im
(
S
(
δ
t
⊥
)
)
=
0
{\displaystyle {\text{Re}}(S(\delta t_{\perp }))={\text{Im}}(S(\delta t_{\perp }))=0}
then
0
≥
1
−
2
π
E
a
v
g
δ
t
⊥
ℏ
{\displaystyle 0\geq 1-{\frac {2}{\pi }}{\frac {E_{avg}\delta t_{\perp }}{\hbar }}}
.
For this bound to become an equality we demand
cos
(
x
)
=
1
−
2
π
(
x
+
sin
(
x
)
)
{\displaystyle \cos(x)=1-{\frac {2}{\pi }}(x+\sin(x))}
, that is
x
=
0
{\displaystyle x=0}
or
x
=
π
{\displaystyle x=\pi }
. Thus
E
n
δ
t
⊥
ℏ
=
0
or
E
n
δ
t
⊥
ℏ
=
π
∀
n
,
c
n
≠
0
{\displaystyle {\frac {E_{n}\delta t_{\perp }}{\hbar }}=0\quad {\text{or}}\quad {\frac {E_{n}\delta t_{\perp }}{\hbar }}=\pi \quad \forall n,c_{n}\neq 0}
,
which holds for only two energy eigenstates
E
0
=
0
{\displaystyle E_{0}=0}
and
E
1
=
π
ℏ
δ
t
⊥
{\displaystyle E_{1}={\frac {\pi \hbar }{\delta t_{\perp }}}}
. Thus, the only state that attains this bound is a two-level pure quantum state (qubit ) in an equal superposition
|
ψ
q
⟩
=
1
2
(
e
i
φ
0
|
0
⟩
+
e
i
φ
1
|
π
ℏ
δ
t
⊥
⟩
)
{\displaystyle \left|\psi _{q}\right\rangle ={\frac {1}{\sqrt {2}}}\left(e^{i\varphi _{0}}\left|0\right\rangle +e^{i\varphi _{1}}\left|{\frac {\pi \hbar }{\delta t_{\perp }}}\right\rangle \right)}
of energy eigenstates
|
E
0
⟩
{\displaystyle \left|E_{0}\right\rangle }
and
|
E
1
⟩
{\displaystyle \left|E_{1}\right\rangle }
, unique up to degeneracy of the energy level
E
1
{\displaystyle E_{1}}
and arbitrary phase factors
φ
0
{\displaystyle \varphi _{0}}
,
φ
1
{\displaystyle \varphi _{1}}
of the eigenstates.[ 2]
Time-varying Hamiltonian
The Margolus-Levitin theorem generalizes to the case with time-varying Hamiltonian and mixed states.[ 5]
Let
H
δ
t
{\displaystyle H_{\delta t}}
be the Hamiltonian at time interval
δ
t
{\displaystyle \delta t}
, such that
H
δ
t
{\displaystyle H_{\delta t}}
still has zero energy in the ground state. Let the system start at some mixed state with density operator
ρ
0
{\displaystyle \rho _{0}}
and evolve by the Schrödinger equation over time. Then
∫
0
δ
t
|
t
r
(
ρ
0
H
δ
t
)
|
d
t
≥
ℏ
D
B
(
ρ
0
,
ρ
δ
t
)
{\displaystyle \int _{0}^{\delta t}|tr(\rho _{0}H_{\delta t})|dt\geq \hbar D_{B}(\rho _{0},\rho _{\delta t})}
,
where
D
B
{\displaystyle D_{B}}
is the Bures distance between the starting state and the ending state.
To obtain the original theorem, set
H
δ
t
{\displaystyle H_{\delta t}}
to be independent of time, and
ρ
0
=
|
ψ
(
0
)
⟩
⟨
ψ
(
0
)
|
{\displaystyle \rho _{0}=\left|\psi (0)\right\rangle \left\langle \psi (0)\right|}
, then since pure states evolve to pure states,
ρ
δ
t
=
|
ψ
(
δ
t
)
⟩
⟨
ψ
(
δ
t
)
|
{\displaystyle \rho _{\delta t}=\left|\psi (\delta t)\right\rangle \left\langle \psi (\delta t)\right|}
, and so by the formula for the Bures distance between pure states ,
E
a
v
g
δ
t
≥
ℏ
arccos
|
⟨
ψ
(
0
)
|
ψ
(
δ
t
)
⟩
|
{\displaystyle E_{avg}\delta t\geq \hbar \arccos |\left\langle \psi (0)|\psi (\delta t)\right\rangle |}
,
and when the starting and ending states are orthogonal, we obtain
E
a
v
g
δ
t
⊥
≥
ℏ
π
2
{\displaystyle E_{avg}\delta t_{\perp }\geq \hbar {\frac {\pi }{2}}}
.
However, the Margolus–Levitin theorem has not yet been established in time-dependent quantum systems, whose Hamiltonians
H
δ
t
{\displaystyle H_{\delta t}}
are driven by arbitrary time-dependent parameters, except for the adiabatic case.[ 6]
Other relevant theorems
Relevant theorems concerning the Margolus–Levitin and the Mandelstam-Tamm theorems were proved[ 2] in 2009 by Lev B. Levitin and Tommaso Toffoli .
Theorem
In the case
E
a
v
g
≠
δ
E
{\displaystyle E_{avg}\neq \delta E}
the orthogonalization interval satisfies
δ
t
⊥
≤
π
ℏ
(
1
+
e
ln
|
δ
E
E
a
v
g
|
)
2
E
a
v
g
(
1
+
δ
E
E
a
v
g
)
(
1
+
ϵ
)
=
π
ℏ
2
E
a
v
g
(
1
+
ϵ
)
,
∀
ϵ
>
0
{\displaystyle \delta t_{\perp }\leq {\frac {\pi \hbar \left(1+e^{\ln {\left|{\frac {\delta E}{E_{avg}}}\right|}}\right)}{2E_{avg}\left(1+{\frac {\delta E}{E_{avg}}}\right)}}\left(1+\epsilon \right)={\frac {\pi \hbar }{2E_{avg}}}\left(1+\epsilon \right),\quad \forall \epsilon >0}
Theorem
For any state
|
ψ
⟩
{\displaystyle \left|\psi \right\rangle }
E
m
a
x
4
≤
E
a
v
g
≤
E
m
a
x
2
{\displaystyle {\frac {E_{max}}{4}}\leq E_{avg}\leq {\frac {E_{max}}{2}}}
,
where
E
m
a
x
{\displaystyle E_{max}}
is the maximum energy eigenvalue of
|
ψ
⟩
{\displaystyle \left|\psi \right\rangle }
and
π
ℏ
≤
E
m
a
x
δ
t
⊥
≤
2
π
ℏ
{\displaystyle \pi \hbar \leq E_{max}\delta t_{\perp }\leq 2\pi \hbar }
,
wherein
E
m
a
x
δ
t
⊥
=
π
ℏ
{\displaystyle E_{max}\delta t_{\perp }=\pi \hbar }
for the qubit state
|
ψ
q
⟩
{\displaystyle \left|\psi _{q}\right\rangle }
with
E
1
=
E
m
a
x
{\displaystyle E_{1}=E_{max}}
.
Proof
Let
S
(
δ
t
⊥
)
=
⟨
ψ
(
0
)
|
ψ
(
δ
t
⊥
)
⟩
=
∑
n
|
c
n
|
2
e
−
i
δ
t
⊥
ℏ
E
n
=
0
{\displaystyle S(\delta t_{\perp })=\left\langle \psi (0)|\psi (\delta t_{\perp })\right\rangle =\sum _{n}|c_{n}|^{2}e^{-i{\frac {\delta t_{\perp }}{\hbar }}E_{n}}=0}
.
Assume a contrario that
E
m
a
x
>
2
π
ℏ
δ
t
⊥
{\displaystyle E_{max}>{\frac {2\pi \hbar }{\delta t_{\perp }}}}
. We can define
E
l
≐
E
m
a
x
−
2
π
ℏ
δ
t
⊥
>
0
{\displaystyle E_{l}\doteq E_{max}-{\frac {2\pi \hbar }{\delta t_{\perp }}}>0}
. But then
e
−
i
δ
t
⊥
ℏ
E
l
=
e
−
i
δ
t
⊥
ℏ
E
m
a
x
e
2
π
i
=
e
−
i
δ
t
⊥
ℏ
E
m
a
x
{\displaystyle e^{-i{\frac {\delta t_{\perp }}{\hbar }}E_{l}}=e^{-i{\frac {\delta t_{\perp }}{\hbar }}E_{max}}e^{2\pi i}=e^{-i{\frac {\delta t_{\perp }}{\hbar }}E_{max}}}
.
Thus, replacing
E
m
a
x
{\displaystyle E_{max}}
with
E
l
>
E
m
a
x
{\displaystyle E_{l}>E_{max}}
does not change
S
(
δ
t
⊥
)
{\displaystyle S(\delta t_{\perp })}
and therefore the set of energy eigenvalues is bounded from above.[ 2] To prove the existence of the lower bound on
E
m
a
x
{\displaystyle E_{max}}
, let the average energy be
E
a
v
g
(
1
)
{\displaystyle E_{avg}^{(1)}}
. We note that replacing energy levels
E
n
{\displaystyle E_{n}}
in
S
(
δ
t
⊥
)
{\displaystyle S(\delta t_{\perp })}
with
E
m
a
x
−
E
n
{\displaystyle E_{max}-E_{n}}
will not affect its validity. But after such a replacement, the average energy is
E
a
v
g
(
2
)
=
E
m
a
x
−
E
a
v
g
(
1
)
{\displaystyle E_{avg}^{(2)}=E_{max}-E_{avg}^{(1)}}
, and we can choose
E
a
v
g
=
min
(
E
a
v
g
(
1
)
,
E
a
v
g
(
2
)
)
{\displaystyle E_{avg}=\min \left(E_{avg}^{(1)},E_{avg}^{(2)}\right)}
. Thus
E
a
v
g
≤
E
m
a
x
2
{\displaystyle E_{avg}\leq {\frac {E_{max}}{2}}}
. Using the bound on
E
a
v
g
{\displaystyle E_{avg}}
from the Margolus–Levitin theorem completes the proof.[ 2]
Furthermore, if
δ
t
⊥
=
π
ℏ
E
m
a
x
{\displaystyle \delta t_{\perp }={\frac {\pi \hbar }{E_{max}}}}
then
S
(
δ
t
⊥
)
=
∑
n
=
0
m
|
c
n
|
2
e
−
i
π
E
n
E
m
a
x
=
∑
n
=
0
m
|
c
n
|
2
(
cos
(
π
E
n
E
m
a
x
)
−
i
sin
(
π
E
n
E
m
a
x
)
)
=
0
{\displaystyle S(\delta t_{\perp })=\sum _{n=0}^{m}|c_{n}|^{2}e^{-i\pi {\frac {E_{n}}{E_{max}}}}=\sum _{n=0}^{m}|c_{n}|^{2}\left(\cos \left(\pi {\frac {E_{n}}{E_{max}}}\right)-i\sin \left(\pi {\frac {E_{n}}{E_{max}}}\right)\right)=0}
,
which is satisfied[ 2] iff
E
0
=
0
{\displaystyle E_{0}=0}
,
E
1
=
E
m
a
x
=
π
ℏ
δ
t
⊥
{\displaystyle E_{1}=E_{max}={\frac {\pi \hbar }{\delta t_{\perp }}}}
, and
|
c
n
|
2
=
1
2
{\displaystyle |c_{n}|^{2}={\frac {1}{2}}}
.
See also
References
^ a b
Leonid Mandelstam; Igor Tamm (1945), "The Uncertainty Relation Between Energy and Time in Non-relativistic Quantum Mechanics", J. Phys. (USSR) , 9 : 249–254
^ a b c d e f g h i
Lev B. Levitin; Tommaso Toffoli (2009), "Fundamental Limit on the Rate of Quantum Dynamics: The Unified Bound Is Tight" , Physical Review Letters , 103 (16): 160502, arXiv :0905.3417 , Bibcode :2009PhRvL.103p0502L , doi :10.1103/PhysRevLett.103.160502 , ISSN 0031-9007 , PMID 19905679 , S2CID 36320152
^ Yakir Aharonov; Jeeva Anandan (1990), "Geometry of quantum evolution", Physical Review Letters , 65 (14): 1697–1700, Bibcode :1990PhRvL..65.1697A , doi :10.1103/PhysRevLett.65.1697 , PMID 10042340
^
Norman Margolus; Lev B. Levitin (1998), "The maximum speed of dynamical evolution", Physica D , 120 (1–2): 188–195, arXiv :quant-ph/9710043 , Bibcode :1998PhyD..120..188M , doi :10.1016/S0167-2789(98)00054-2 , S2CID 468290
^
Deffner, Sebastian; Lutz, Eric (2013-08-23). "Energy–time uncertainty relation for driven quantum systems" . Journal of Physics A: Mathematical and Theoretical . 46 (33): 335302. arXiv :1104.5104 . Bibcode :2013JPhA...46G5302D . doi :10.1088/1751-8113/46/33/335302 . hdl :11603/19394 . ISSN 1751-8113 . S2CID 119313370 .
^
Okuyama, Manaka; Ohzeki, Masayuki (2018). "Comment on 'Energy-time uncertainty relation for driven quantum systems' " . Journal of Physics A: Mathematical and Theoretical . 51 : 318001. arXiv :1802.00995 . doi :10.1088/1751-8121/aacb90 . ISSN 1751-8113 .
Deffner, Sebastian; Campbell, Steve (2017), "Quantum speed limits", Journal of Physics A , 50 (45): 453001, arXiv :1705.08023 , Bibcode :2017JPhA...50S3001D , doi :10.1088/1751-8121/aa86c6 , S2CID 3477317
Jordan, Stephen P. (2017), "Fast quantum computation at arbitrarily low energy", Physical Review A , 95 (3): 032305, arXiv :1701.01175 , Bibcode :2017PhRvA..95c2305J , doi :10.1103/PhysRevA.95.032305 , S2CID 118953874
Lloyd, Seth ; Ng, Y. Jack, "Black Hole Computers ", Scientific American (April 2007), p. 53–61
Sinitsyn, Nikolai A. (2018). "Is there a quantum limit on speed of computation?". Physics Letters A . 382 (7): 477–481. arXiv :1701.05550 . Bibcode :2018PhLA..382..477S . doi :10.1016/j.physleta.2017.12.042 . S2CID 55887738 .