ArticlePDF Available

Decoherence-free dynamical and geometrical entangling phase gates

Authors:

Abstract and Figures

It is shown that entangling two-qubit phase gates for quantum computation with atoms inside a resonant optical cavity can be generated via common laser addressing, essentially, within one step. The obtained dynamical or geometrical phases are produced by an evolution that is robust against dissipation in form of spontaneous emission from the atoms and the cavity and demonstrates resilience against fluctuations of control parameters. This is achieved by using the setup introduced by Pachos and Walther [Phys. Rev. Lett. 89, 187903 (2002)] and employing entangling Raman- or STIRAP-like transitions that restrict the time evolution of the system onto stable ground states. Comment: 10 pages, 9 figures, REVTEX, Eq. (20) corrected
Content may be subject to copyright.
arXiv:quant-ph/0309180v3 28 Aug 2005
Decoherence-free dynamical and geometrical entangling phase gates
Jiannis K. Pachos
and Almut Beige
Blackett Laboratory, Imperial College London, Prince Consort Road, London, SW7 2BW, UK
(February 1, 2008)
It is shown that entangling two-qubit phase gates for quantum computation with atoms inside
a resonant optical cavity can be generated via common laser addressing, essentially, within one
step. The obtained dynamical or geometrical phases are produced by an evolution that is robust
against dissipation in form of spontaneous emission from the atoms and the cavity and demonstrates
resilience against fluctuations of control parameters. This is achieved by using the setup introduced
by Pachos and Walther [Phys. Rev. Lett. 89, 187903 (2002)] and employing entangling Raman- or
STIRAP-like transitions that restrict the time evolution of the sy stem onto stable ground states.
03.67.Pp, 42.50.Pq
I. INTRODUCTION
The phase of a state vector is undoubtedly one of the
central curiosities that differentiate quantum from clas-
sical mechanics. Especially, the presence of quantum
phases in non-intuitive interference expe riments has been
the focus of intellectual excitement in many studies (see
for example [1–5]). The presence of a phase is e ven more
striking in a state that exhibits entanglement. In par -
ticular, the research on highly correlated subsystems has
contributed to the boosting of innovative technological
applications [6–8]. Such an example is quantum compu-
tation where the entangling phase gate often provides the
basic two-qubit gate an essential ingredient for the real-
ization of arbitrary quantum algorithms. This gate cor-
responds to a unitary oper ation that changes the phase
of one subsystem depending on the state of another one.
In the last years, considerable effort has been made to
find efficient ways for its experimental implementation.
For a proposed gate implementa tion to be feasible with
present technology, it is importa nt that the scheme is
widely independent from various control parameters like
the operation time. Minimizing the control errors will
augment the efforts for engineering scalable quantum in-
formation processing with high accuracy. Another prob-
lem arises from the fact that it is very difficult to isolate
a quantum mechanical system c ompletely from its en-
vironment without loosing the possibility to manipulate
its state. Uncontrollable enviro nmental couplings lead
in general to dissipation and the loss of information. In
addition, the phase of a state vector is very fragile with
respect to decoherence. As a result, phase factors are
rather hard to generate and to store in engineered sys-
tems.
This paper analyzes in detail entangling phase oper-
ations that are especially designed to bypass the dissi-
pation problem and guar antee very high fidelities for a
wide range of experimental parameters. In particular, we
consider the quantum behaviour of atoms inside an op-
tical cavity which has been observed experimentally by
Hennrich et al. [9] and, more recently, by J. McKeever et
al. [10] and by Sauer et al. [11 ]. The scheme presented
here involves two atoms trapped at fixed positions inside
an optical cavity and can be implemente d using the tech-
nology of the recent calcium ion experiment [12]. Alter-
natively, neutral atoms can be tr app e d with a standing
laser field [13], as in the exp e riment by Fischer et al. [14],
or in an optical lattice [11 ,15]. In the last decade, several
atom-cavity quantum computing schemes have been pro-
posed [16–24], e ach of them having its respec tive merits.
To implement quantum phase gates we utilize adia-
batic processes that result in the generation of dynam-
ical or geometrical phases [25]. Employing geometrical
phases is an intriguing way of manipulating quantum me-
chanical systems. They exhibit independence from the
operation time of the control evolution and are robust
against per tur bations of system parameters as long as
certain requirements, concerning the geometrical charac-
teristics of the evolution, a re satisfied. To avoid dissipa-
tion, we exploit the existence of dec oherence-free states
[26–29]. As in [20], only ground states with no pho-
ton in the cavity mode and the atoms in a stable state
become populated. Population transfers between those
ground states are achieved with the help of Raman or
STIRAP-like processes [30,31]. Their control pr ocedures
are ex actly the same a s for the usual Raman and STI-
RAP transitions, but now the c oupling of the atoms to
the same cavity mode allows for the creation of entan-
glement. Another advantage of the proposed scheme is
that it can be realized via common laser addressing and,
essentially, within one step.
In the following, each qubit is obtained from two
ground states of the same atom. High fidelities of the
final state are achieved even for moderate values o f the
jiannis.pachos@imperial.ac.uk
1
atom-cavity c oupling co nstant g compared to the spon-
taneous cavity decay rate κ and the atom decay rate
Γ. Different fro m other atom-cavity quantum comput-
ing schemes [16–19,21], we assume the c onstant g
2
to be
about 100 κΓ which is close to the state of the art technol-
ogy [11] and should be within the range of experiments
in the nearer future [15].
The paper is organized as follows. Section II intro-
duces the decoherence-free subspace of the sy stem with
respect to leakage of photons through the cavity mir-
rors. An effective Hamiltonian describing the possible
time evolutions of the system within that subspace is
derived. In order to avoid spontaneous emission from
the atoms, we present in Section III different parame-
ter regimes that minimize the population in the excited
atomic levels and result in entangling Raman and entan-
gling STIRAP transitions. In Section I V we employ the
correspo nding time evolutions to generate dynamical and
geometrical phases for the rea lization of two-qubit phase
gates for quantum computation. Finally, we s ummarize
our results in Section V.
II. ELIMINATION OF CAVITY DECAY
In the following we consider two ato ms (or ions), each
of them comprising a four-level system. A possible level
configuration, which is suited for the implementation of
quantum computing with calcium ions, is shown in Fig-
ure 1. Each qubit is obtained from the stable ground
states |0i and |1i of the same atom. In addition, each
atom posses ses an e xcited state |2i, that becomes only
virtually populated during gate operations, and a third
auxiliary ground state |σi. In order to realize an entan-
gling opera tion, the two atoms involved should be posi-
tioned in the cavity, where both see the same coupling
constant g, and laser fields should be applied as shown.
As we see later, the detuning δ plays an important role
in regulating the phase factor of the final state of the
system.
atom 2:
σ
+
σ
qubit 1
qubit 2
2
σ
0
1
g
S
1/2
3/2
D
1/2
P2
σ
0
1
g
S
1/2
3/2
D
1/2
P
atom 1:
∆−δ
∆−δ
+
σ
σ
FIG. 1. Level configuration for atom-cavity quantum com-
puting with calcium ions showing the d riv ing laser fields. The
2-σ and the 2-1 transitions are activated by the laser radia-
tions with amplitudes and polarizations given by
σ
, σ
+
and
1
, σ
respectively. Each qubit is obtained from two degen-
erate S
1/2
ground states of the same ion.
Let us denote the cavity photon a nnihilatio n and cre-
ation operator by b and b
while is the detuning of the
cavity with respect to the 2-1 transition of each atom.
The occupation of the cavity is then given by n = hb
bi.
One laser drives the 2-1 transition of atom i with Rabi
frequency
(i)
1
and detuning ∆; another one excites the
2-σ transition with Rabi frequency
(i)
σ
and detuning
δ. Going over to the interaction picture with re-
sp e c t to H
0
+ ¯h
P
i
[ δ|σi
ii
hσ| + |2i
ii
h2|] where H
0
is
the interaction-free Hamiltonian, one finds
H
I
=
2
X
i=1
¯hg
|2i
ii
h1|b + H.c.
+
1
2
¯h
(i)
1
|1i
ii
h2| +
(i)
σ
|σi
ii
h2| + H.c.
¯ |σi
ii
hσ| ¯h |2i
ii
h2| . (1)
The generatio n of a non-trivial time evolution requires
in gener al a non-homoge neity in the Rabi frequencies of
the laser fields with respect to the two atoms. Without
loss of generality, we assume here that the Rabi frequen-
cies of laser fields coupling to the same transition differ
only in phase. In the fo llowing, e ntangling operations are
realized by choosing
(1)
σ
=
(2)
σ
σ
,
(1)
1
=
(2)
1
1
/
2 . (2)
These parameters can be implemented via common laser
addressing by driving each transition with the same laser
field and from an angle that produces a π phase difference
between atom 1 and 2.
One of the main sources for decoherence in atom-cavity
setups is the leakage of photons through the cavity mir-
rors. However, the presence of high spontaneous de-
cay rates does not necessar ily lead to dissipation. As it
has been shown in the past, the presence of a relatively
strong atom-cavity co upling constant g [32] c an have
the same effect as continuous measurements whether the
cavity mode is empty or not. As a consequence, the
time evolution of the system becomes restricted onto the
decoherence-free states of the system with respect to cav-
ity decay. In this Section, we exploit this effect and its
consequences for the time evolution to significantly sim-
plify the requirements for atom-cavity quantum compu-
tation.
In the next Section, Raman and STIRAP-like pro-
cesses [30,31] are introduced in order to minimize the
population in the excited atomic state |2i. Assuming
that the population in level 2 is negligible, the only con-
dition that has to be fulfilled to avoid cavity decay is
that the Rabi freq uencies
1
and
σ
are sufficiently weak
compared to the atom-cavity coupling constant g,
|
1
|, |
σ
| g (3)
This relation induces two different time scales in the sys-
tem and allows for the calculation of the time evolution
with the help of an adiabatic elimination. Figure 2 shows
2
the most r e le va nt transitions if the sy stem is initially pre-
pared in a qubit state with an empty cavity (n = 0) and
uses the abbreviations
|αi
|12i |21i
/
2 ,
|Ai
|σ1i + |1σi
/
2 ,
|˜αi
|σ2i |2σi
/
2 ,
|
˜
Ai
|σ1i |1σi
/
2 . (4)
Setting the amplitudes of all fast varying s tates equal to
zero and using the Hamiltonian (1), one finds that the
system evolves effectively according to the Hamiltonian
H
eff
=
1
2
¯h
1
|αih11| +
σ
|αihA| + H.c.
¯ |AihA| ¯h |αihα|
|0
cav
ih0
cav
| . (5)
As expected for adiabatic processes, this Hamiltonian re-
stricts the time evolution onto a subspace of slowly vary-
ing states, which is why the states |˜αi and |
˜
Ai are not
present in Equation (5). The subspace of the slowly vary-
ing states includes, in addition to the ground sta tes , only
one state w ith population in the excited state |2i. This
state, |αi|0
cav
i, is a zero eigenstate of the atom-cavity
interaction.
...
g
...
g
laser excitation
to other levels via
01,n=0
02,n=0
10,n=0
20,n=001,n=1
...
1
1
10,n=1
11,n=0
α,n=0
σσ,
n=0
α,n=0
~
A,n=0
...
g
1
1
σ
laser excitation
to other levels via
σ
A,n=1
~
∆−δ
∆−δ
FIG. 2. Level configuration showing the most relevant
transitions for the initial qub it states |11i, |01i and |10i in
the presence of mechanisms (see Section III) that minimize
the population in level 2. The presence of a strong cavity cou-
pling constant g prohibits a transfer of population between
|Ai and |σσi. This does not effect t he t ransition between |11i
and |Ai. The atomic state |00i does not see any laser fields
and does not change its amplitude in time.
Note that the Hamiltonian (5) indeed restricts the time
evolution of the sy stem onto its deco herence-free sub-
space with respec t to cavity decay. A state belongs to
this subspace if the resonator mode is in its vacuum state
|0
cav
i and the atom-cavity interaction cannot transfer ex-
citation from the atoms into the resonator. Hence, the
decoherence-free s ubspace includes all gr ound states and
the state |αi|0
cav
i. They are exactly the slowly varying
states of the system. Furthermore, the effective Ha mil-
tonian can be written as
H
eff
= IP
DFS
H
laser
IP
DFS
, (6)
where IP
DFS
is the projector onto the dec oherence-free
subspace
IP
DFS
=
h
|αihα| +
X
i,j=0,1
|ijihij|
i
|0
cav
ih0
cav
| , (7)
and H
laser
is the laser term in (1). One way to interpret
this is to state that condition (3) effectively induces con-
tinuous measurements whether the cavity field is empty
or not [33]. As a consequence, entanglement can be gen-
erated between qubits via excitation of the ma ximally
entangled state |αi.
In the event that condition (3) is not fulfilled, as in all
realistic experiments, the ab ove argumentation does not
hold and corrections to the effective time evolution (5)
have to be taken into acc ount. To identify the evolution
of the system more accurately, we use the quantum jump
approach [34–36] which predicts the no-photon time evo-
lution of the sy stem with the help of the conditional, non-
Hermitian Hamiltonian H
cond
. Given the initial state
|ψi, the state of the system equals U
cond
(T, 0) |ψi/k·k at
time T under the condition of no emission. For conve-
nience, H
cond
has been defined such that
P
0
(T ) = kU
cond
(T, 0) |ψik
2
(8)
is the probability for no photon in (0, T ). If photo ns
are emitted, then the computation failed and has to be
repeated. To some extent, this can be compensated by
monitoring photon emissions with good detectors. Alter-
natively, quantum teleportation can be employed to per-
form a whole algorithm by selecting the successful gates
[37].
Let us now consider a concrete example to see how
well the elimination of cavity decay works. For the two
four-level atoms in Figure 1 it is [20]
H
cond
= H
I
i
2
¯ b
b
i
2
¯hΓ
2
X
i=1
|2i
ii
h2| . (9)
Suppose that the sy stem is initially prepared in |01i|0
cav
i
and a laser field is applied with
1
6= 0 that excites the
2-1 transition of both atoms for a significant time T .
Equation (5) then implies the inhibition of any time evo-
lution since |01i|0
cav
i is a zero eigenstate of H
eff
. Indeed,
the numerical solution of the time evolution given by the
Hamiltonian (9) reveals that the final state coincides with
the initial state of the system with fidelity F 1 under
the condition of no photon emission in (0, T ). The suc-
cess rate P
0
(T ) as a function of and
1
is shown in
Figure 3. As expected, P
0
is the closer to one the smaller
3
the Rabi fr e quency
1
and its size is widely independent
from the detuning ∆.
FIG. 3. The success rate P
0
(T ) as a function of the
Rabi frequency
1
and the detuning for the initial state
|01i|0
cav
i, κ = Γ = 0.1 g, δ = 0 and T = 2000/g. If one al-
lows for a short transition time at the end of the operation for
unwanted states to decay, then the fin al state of the system
equals exactly the initial one.
III. ELIMINATION OF ATOM DECAY AND
ENTANGLING OPERATIONS
For optical cavities, the atom decay rate Γ is in gen-
eral of about the s ame size as the parameters g and κ.
We therefore still need to overcome the problem of spon-
taneous emission from the atoms in order to realize a
coherent time evolution. This can be achieved by mini-
mizing the population in level 2 and keeping the system
effectively in the ground states |11i|0
cav
i and |Ai|0
cav
i.
To realize this we employ in the following Raman or
STIRAP-like transitions [30,31]. These are directly ap-
plicable since the relevant states comprise a Λ-type three-
level configuration. Entanglement between the atoms is
created by coupling |11i to the maximally entangled state
|Ai.
A. Entangling Raman transitions
One way to implement transitions without populating
the excited state |αi is to choose the detuning much
larger than the Rabi frequencies
1
and
σ
,
|
1
|, |
σ
| (10)
This choice of parameters results in the realization of a
Raman-like transition since the la rge detuning intro-
duces an additional time scale in the time e volution (5).
Adiabatically eliminating the excited state |αi reveals
that the system is e ffectively g overned by the Hamilto-
nian
H
eff
=
1
2
¯h
|11ihA| + H.c.
¯h
11
|11ih11| ¯h
A
|AihA| (11)
with the abbreviations
1
σ
2∆
,
11
|
1
|
2
4∆
,
A
δ
|
σ
|
2
4∆
. (12)
Solving the time evolution for the case where the lasers
are turned on for a period T with constant amplitude
yields
U
eff
(T, 0) = exp
i
2
(∆
11
+
A
)T
h
cos
KT
2
+
i(∆
11
A
)
K
sin
KT
2
|11ih11|
K
sin
KT
2
|11ihA| + |Aih11|
+
cos
KT
2
i(∆
11
A
)
K
sin
KT
2
|AihA|
i
, (13)
where
K
||
2
+ (∆
11
A
)
2
1/2
(14)
plays the role of a Ra bi frequency. As the evolution (13)
can be used to create entanglement, we call it an entan-
gling Raman (E-Raman) transition.
In order to see how well the elimination of the dissi-
pative states works, let us consider as an example the
preparation of the maximally entangled s tate |Ai. If the
atoms are initially in |11i, this can be achieved by apply-
ing a laser pulses of length T = π/K = 2π|
1
|
2
with
δ = 0 and
σ
=
1
. Figures 4 and 5 present the fidelity
F and success rate P
0
, respec tively, of this entangling
operation as a function of the Rabi frequency
1
and the
detuning ∆. They result from a numerical integration of
the time evolution with the full Hamiltonian (9).
In particular for κ = Γ = 0.1 g, the fidelity of the final
state can be as high as 0.993 if one chooses
1
= 0.01 g
and = 1.357 g and allows the success probability to
be as high as P
0
= 0.857. Higher fidelities can only be
obtained in the presence of sma ller decay rates. In gen-
eral one can improve the success rate P
0
by a few percent
by increasing the detuning and by sacrificing an e rror
from the maximal fidelity of the order of 10
3
. While
the fidelity exhibits a maximum as a function of ∆, the
success probability increases monotonically with ∆. For
large detunings, the adiabatic elimination of the cavity
mode is no longer efficient (see Figure 2) resulting in the
reduction of F . Varying the Rabi freq uency gives larger
fidelities for smaller
1
since the adiabatic elimination
4
of cavity excitation as well as the elimination of |αi be-
comes mo re efficient. However, the smaller
1
the la rger
the operation time T of the entangling evo lution.
FIG. 4. The fidelity F for the preparation of the maximally
entangled state |Ai given the initial state |11i as functions of
the Rabi frequency
1
and the detuning for
σ
=
1
,
Γ = κ = 0.1 g and δ = 0.
FIG. 5. The success rate P
0
for the preparation of the
maximally entangled state |Ai for the same parameters as
in Figure 4.
B. Entangling STIRAP processes
Another way to transfer population between the
ground states |11i and |Ai without ever populating the
excited state |αi is to employ a STIRAP-like process
[30,31]. As this is an entangling process we can call it
an entangling STIRAP (E- STIRAP ) transition. To see
how the s uppression of spontaneous emission from the
atoms works in this case let us consider again the ef-
fective Hamiltonian (5) and note that its eigenvalues for
δ are
E
0
= 0 , E
±
=
1
2
( ±
p
|
σ
|
2
+ |
1
|
2
+
2
) . (1 5)
For convenience, we now introduce the interaction pic-
ture with respect to the Hamiltonian H
0
= ¯|AihA|.
Within that picture, the eigenvectors of the effective
Hamiltonian become time dependent and the dark state
of the system equals
|E
0
i
1
p
|
σ
|
2
+ |
1
|
2
σ
|11i e
iδt
1
|Ai
. (16)
The other two eigenstates |E
±
i occupy the antisymmet-
ric state |αi and hence possess population in the excited
atomic state |2i.
Spontaneous emission fr om the atoms can therefore be
avoided if the s ystem remains constantly in the dark state
|E
0
i. Nevertheless, a nontrivial entangling time evolution
between the gr ound states |11i and |Ai can be induced,
by va rying the laser amplitudes
1
and
σ
in time. If
this happens slow c ompared to the time scale given by
the eigenvalues E
±
, no population is transferred into the
eigenstates |E
±
i. Assuming adiabaticity, the system fol-
lows the changing parameters and remains in the dark
state. The fact that |E
0
i incorporates a time dependent
phase factor can be used to induce a certain phase on
the final state |Ai by choosing the small laser detuning δ
appropriately.
In the following we assume = 0 since this maxi-
mizes the energy distance between |E
0
i and its nearest
eigenstate. For convenience, we also consider the case
where
1
and
σ
are real and define tan θ
1
/
σ
and
φ δt. The eigenstate corresponding to the zero eigen-
value then take s the familiar for m
|E
0
i = co s θ |11i e
iφ
sin θ |Ai . (17)
The adiabatic condition is satisfied if
|
˙
θ|, |
˙
φ|
p
2
σ
+
2
1
(18)
If the time dependence of θ and φ is indeed such that
the initial state of the system coincides with the dark
state |E
0
(0)i, the effective evolution along a path C in
the (θ, φ) parameter space can easily be calculated. As a
consequence of the adiabatic theorem, the effective evo-
lution is diagonal with respect to the basis given by the
eigenvectors |E
0
i and |E
±
i. Especially for the case where
θ(0) = 0 and with respect to the basis states |E
0
(0)i,
|E
+
(0)i and |E
(0)i, the effective time evolution of the
system can be wr itten as
U
eff
(T, 0) = R(θ, φ)
e
iϕ
0
0 0
0 e
iϕ
+
0
0 0 e
iϕ
(19)
with θ = θ(T ), φ = φ(T ) and
R(θ, φ) =
1
2
×
2 cos θ
2 sin θ
2 sin θ
2e
iφ
sin θ (1 + e
iφ
cos θ) (1 e
iφ
cos θ)
2e
iφ
sin θ (1 e
iφ
cos θ) (1 + e
iφ
cos θ)
. (20)
5
The phases ϕ
i
ϕ
d
i
+ ϕ
g
i
(i = 0, ±) are in ge neral the
sum of a dynamical and a geometrical phase with
ϕ
d
0
= 0 , ϕ
d
+
= ϕ
d
=
1
2
Z
T
0
q
2
1
+
2
σ
dt (21)
while the geometrical phases equal
ϕ
g
0
=
I
C
sin
2
θ dφ , ϕ
g
+
= ϕ
g
=
1
2
I
C
cos
2
θ dφ (22)
for a closed circular loop C.
As an example, let us consider the case where the ini-
tial state |11i is transferred into the maximally entangled
state |Ai. This can be achieved by varying the Rabi fre-
quencies
1
and
σ
slowly in a so-called counterintuitive
pulse sequence. Figure 6 and 7 show the result of a nu-
merical solution of the corresponding time evolution tak-
ing the full Hamiltonian (9) into account. In particula r,
we calculated the fidelity and the success of the scheme
as a function of the maximum laser amplitude and the
frequency ω for = δ = 0, κ = Γ = 0.1 g and
σ
(t) =
(
sin ωt , for t
0,
2
3
T
0 , for t
2
3
T, T
(23)
and
1
(t) =
(
0 , for t
0,
1
3
T
sin ω
t
1
3
T
, for t
1
3
T, T
(24)
with the total operation time T = 3π/(2ω).
FIG. 6. The fidelity F for the preparation of the maximally
entangled state |Ai as a function of the maximum Rabi fre-
quency and frequency ω for = δ = 0 and κ = Γ = 0.1 g.
FIG. 7. The success rate P
0
for the proposed entangled
state preparation scheme for the same parameters as in Fig-
ure 6.
In contrast to E-Raman processes, the fidelity F of the
final state can, in principle, b e arbitrarily close to one in-
dependent of the size of the spontaneous decay rates κ
and Γ. For example, for = 0.02 g a nd ω = 4 · 10
5
g
the fidelity becomes F = 0.998 and corresponds to the
success rate P
0
= 0.876 which is still within acceptable
limits. In general, a large leads to the population of
the excited state |2i which might result in the emiss ion of
a photo n w ith rate Γ or in the populatio n of the cavity
mode followed by the leakage of a photon through the
cavity mirror s. As a result, F and P
0
decrease for in-
creasing laser amplitudes Ω. On the other hand, for very
weak Rabi frequencies (Ω 0) the fidelity and the suc-
cess rate increase as ω decreases which results in a longer
operation time T . This is in agreement with condition
(18) which implies that the adiabatic evolution is more
successful for slower evolutions. Neve rtheless, the suc-
cess rate P
0
cannot be increase d arbitrarily. The reason
is that very long oper ation times unavoidably lead to the
population of the cavity mode in the presence of a finite
value for and the elimination of cavity decay does no
longer hold.
IV. DECOHERENCE-FREE DYNAMICAL AND
GEOMETRICAL PHASE GATES
In this Section we employ the decoherence-free E-
Raman and E-STIRAP processes intro duced in the last
Section for the realization of two-qubit quantum gate op-
erations. In particular, we consider the implementation
of the controlled phase gate
CP = dia g
1, 1, 1, e
iϕ
, (25)
where the qubit state |11i collects the phase ϕ while the
other qubit s tates remain unaffected. This gate consti-
tutes, together with a general single-qubit rota tio n, a uni-
versal set of gates and allows for the implementation of
6
any desired unitary time evolution. Attention is paid
to their sta bility a gainst fluctuations of most system pa-
rameters and we analyze in detail the efficiency of various
processes.
A. Raman-based Dynamical Phase Gates
Let us begin with the descr iption of a po ssible imple-
mentation of (25) with the help of the E-Raman process
introduced in Section IIIA. From Equation (13) one sees
that an applied laser pulse of length T evolves the initial
state |11i according to
U
eff
(T, 0)|11 i = exp
i
2
(∆
11
+
A
)T
×
h
cos
KT
2
+ i
11
A
K
sin
KT
2
|11i
K
sin
KT
2
|Ai
i
,
(26)
while the amplitudes of the qubit states |00i, |01i and |10 i
do not change in time. The desired phase gate (25) can
therefore b e realized in the minimal time if one chooses
T = 2π/K . (27)
For this parameter choice, the sine term in (2 6) vanishes
and the cosine term equals 1 giving finally
ϕ = π +
1
2
(∆
11
+
A
)T . (28)
Again, the phase ϕ can be controlled by adjusting the
size of the detuning δ. Especially, for
δ =
1
4∆
|
1
|
2
+ |
σ
|
2
(29)
one obtains ϕ = π a nd the amplitude of the sta te |11i
accumulates a minus sign. This particular evo lution is
illustrated in Figure 8 using a Bloch’s sphere represen-
tation. Note that the condition of timing the evolution
such that the sine term beco mes zero, is robust against
first order timing errors in T due to the behaviour of the
sine function around π.
|11
|11
C
E−Raman
−|11
|A
|A
E−STIRAP
FIG. 8. The E-Raman and E-S TIRAP evolut ions depicted
on t he Bloch sphere. In the E-Raman evolution, the genera-
tion of the phase factor in front of |11i is explicitly shown. I n
the E-STIRAP evolution, the acquired phase relates to the
solid angle spanned by the path C on the sphere.
To minimize the experimental effort for the realiza-
tion of the phase gate (2 5), one should notice that a
non-trivial E-Rama n transition can also be induced using
only one laser field coupling to the 1-2 transition of each
atom. Indeed, for
A
= 0 and the laser pulse duration
T = π/K, Equation (26) yields
U
eff
(T, 0)|11 i = exp
i∆
11
T
|11i . (30)
For this case, Figure 2 shows that the initial state |11 i
couples only to the excited state |αi via the applied heav-
ily detuned las e r field. To see that the pha se operation
(30) is robust against fluctuations of
1
, we now consider
the Rabi frequency being time dependent. In this case
the phase accumulated by the initial state |11i becomes
ϕ = π +
R
T
0
11
dt. This clearly shows that statistical
perturbations of
1
average out due to dependence of
the acquired phase on the integral over
11
along (0, T ).
B. STIRAP-based Geometrical Phase gate
Another possibility to implement the co ntrolled two-
quit phase gate (25) is provided by the E-STIRAP pro-
cess. Let us first consider the case of the g e neration of a
dynamical phase as in [20]. To do so we var y θ from zero
to π, then apply a 2π pulse to transform the state |σi to
−|σi for both atoms and then reduce θ back to zero. At
the end of this time evolution, the state |11i acquires an
overall minus sign, while the other qubit states |00i, |01i
and |10i rema in unchanged. The result is the conditional
phase gate with ϕ = π.
Alternatively, a geometrical phase can be generated by
continuously vary ing the parameters θ and φ in a cyclic
fashion [38,39]. Starting from θ = 0 one can perform a
cyclic adiabatic evolution on the (θ, φ) pla ne along a loop
C. As predicted by Equations (19)-(22), the qubit s tate
|11i then acquires the geometrical phase
ϕ = ϕ
g
0
=
I
C
sin
2
θ dφ (31)
while the dynamical phase ϕ
d
0
is identically zero. The
concrete size of ϕ, which characterizes the phase gate
(25), depends therefore only on the solid angle spanned
by the path C on the Bloch sphere (see Figure 8). Hence
it is very robust against statistical fluctuations of the
Rabi frequencies of the applied laser fields and it is inde-
pendent of the total gate operation time.
Let us now consider a concr e te laser configuration,
where the laser field
σ
is turned on and kept constant.
To perform a non-trivial loop in the control parameter
space (θ, φ) starting from the (0, 0) point we proceed as
follows. As describe d in Section III B, a non-zero detun-
ing δ results in the accumulation of a pha se φ = δt when
7
we introduce a non zero value for θ. Let us consider the
simple case where
θ = arctan (x) with x
1
/
σ
, (32)
changes from zero to a non-zero value. More concretely,
we increase and decrease
1
by linear ramps such that
x(t) =
(
αt , for t
0,
1
2
T
α(T t) , for t
1
2
T, T
(33)
or by a sine r amp given by
x(t) = x
max
sin βt (34)
with t (0, π). For those two cases one can e asily
derive the geometrical phases as functions of the exper-
imental control parameters T , α, x
max
and β and we
present the results in Figure 9.
ϕ/δ
(b)
β
x
max
(a)
ϕ/δ
T
α
FIG. 9. The resulting geometrical phase ϕ
g
produced
(a) by a linear ramp and (b) by a sine ramp. The produ ced
geometrical phases are largely independent of the control pa-
rameters α or x
max
respectively. α and β are depicted in
inverse time units, while x
max
is dimensionless.
From the plots we see that the accumulated geomet-
rical phase ϕ is, as expected, widely independent on the
size of α or x
max
giving the resilience of the proposed E-
STIRAP process from control errors, especially from the
exact value of the laser amplitudes or their time deriva-
tives. This independence of ϕ makes the proposed quan-
tum gate implementation an attractive candidate for ex-
periments where the exact control o f all laser parameters
is not possible. In addition, it is experimentally appealing
to construct a setup where the two employed laser fields
can be produced from the same source, separated in two
by a beam splitter and having the ratio
1
/
σ
modulated
e.g. by an Acousto-Optical Modulator (AOM). With this
setup, the ratio is also resilient to the fluctuations of the
amplitude of the initial laser beam.
V. CONCLUSIONS
In this paper, we discussed in detail concrete proposals
for the realization of universal two-qubit phase gates for
quantum c omputing in atom-cavity setups. The atoms
(or ions) are trapped at fixed positions inside an opti-
cal cavity and each qubit is obtained from two stable
ground states of the same atom. To ac tivate a time evo-
lution, two laser fields a re applied simultaneously induc-
ing transitions in an effectively Λ-like system. In order to
minimize the experimental effort, we seek quantum com-
puting schemes that are as simple as possible in terms of
resources . The quantum gate implementatio ns proposed
here do not require individual addressing of the atoms
and can be realized, essentially, within one step. In par-
ticular, we considered the case where the atom-c avity
coupling constant g is only one order of mag nitude larger
than each of the spontaneous decay rates of the system
and we as sumed g
2
= 100 κΓ.
To avoid dissipation, we populate only the ground state
of the system including all stable ground states o f the
atoms while the cavity is in its vacuum state. Never-
theless, an interaction between qubits can take place via
virtual po pulation of the cavity mode and excited atomic
levels. We first eliminated the possibility of leakage of
photons through the cavity mirrors. This was a chieved
by cho osing the parameters such that adiabaticity re-
stricts the time evolution of the system effectively onto
a deco herence-free subspace with respect to cavity decay
that includes highly entangled atomic states. Using these
states for the implementa tion of Raman and STIRAP-
like evolutions, entanglement can be created betwe e n the
ground states of the atoms.
Comparing the E-Raman and E-STIRAP procedures
we see that they give in general similar fidelities and suc-
cess rates, with E-STIRAP having a slight advantage.
Their control procedures are quite different in bo th c ases,
providing a flexibility to adapt the proposed scheme to
different experimental setups. The main conceptual as
well as practical differe nce appears when we produce the
entangling pha se ga tes . As it is seen in Figure 8 the phase
produced by the E- Raman tra nsition on the state |11i is
given explicitly by the evolution of this state from the
north to the south pole of the Bloch sphere. In contrast,
with the E-STIRAP procedure initial po pulation in the
|11i s tate circula tes along a closed path and returns ex -
actly to the initial point. Nevertheless, at the end the
state |11i acquires a geometrical phase.
The idea of using measurements on ancilla states, like
the measurements on the cavity mode considered in this
paper, leads to a realm of new possibilities for qua ntum
computing. A recent example is the linear optics scheme
by Knill et al. [37] where photons become entangled with-
out ever having to interact. As long as the measurements
on the ancilla state do not reveal any information about
the qubits, they induce a unitary time evolution between
them which can then be used to implement universal gate
operations [40]. In contrast to other schemes, the gate
success rate for atom-cavity quantum computing can, at
least in principle, be arbitrarily close to one since the
measurements are performed almost continuously. This
allows to utilize the quantum Z e no effect [18] which as-
sures that transitions in unwanted states are strongly in-
8
hibited.
The proposed schemes exhibit stability against fluctu-
ations of most system parameters. The presented dy-
namical and geometrical phases are obtained by a time
evolution that does not dependent on the exa c t value
of the e xternal contr ol parameters. Using E-Raman or
E-STIRAP processes, we enjoy the experimental advan-
tages, such as, independence of the final state on the ex-
act intermediate value of the laser field amplitude. Gates
are constructed dynamically as well as geometrically in
order to take advantage of the additional fault-tolerant
features of geometrical quantum computation [41]. Re-
cently, many attempts have been made in exploring the
decoherence cha racteristics of geometrical phases [42,43].
Whether they are advantageous over equivalent dynami-
cal evolutions depends on the employed physical system.
In contrast to this, we presented here evolutions that
are intrinsically decoher e nce -free and naturally allow the
generation of entangling dynamical as well as geometrical
phases with simple control pictures.
Acknowledgements. This work was supported in part
by the Europe an Union and the EPSRC. A.B. thanks
the Royal Society for a James Ellis University Research
Fellowship.
[1] M. O. Scully and K. Dr¨uhl, Phys. Rev. A 25, 2208 (1982);
U. Eichmann, J. C. Bergquist, J. J. Bollinger, J. M. Gilli-
gan, W . M. Itano, D. J. Wineland, and M. G. Raizen,
Phys. Rev. Lett. 70 2359 (1993).
[2] Z. Hradil, R. Myska, J. Perina, M. Zawisky, Y. Hasegawa,
and H. Rauch, Phys. Rev. Lett. 76, 4295 (1996).
[3] D. Bouwmeester, G. P. Karman, C. A. Schrama, J. P.
Woerdman, Phys. Rev. A 53, 985 (1996).
[4] B. Brezger, L. Hackerm¨uller, S. Uttent haler, J.
Petschinka, M. Arndt, and A. Zeilinger, Phys. Rev. Lett.
88, 100404 (2002).
[5] M. Berry, I. Marzoli, and W. Schleich, Phys. World 14,
39 (2001).
[6] J. P. Dowling, Phys. Rev. A 57, 4736 (1998).
[7] J. P. Dowling and G. J. Milburn, Quantum Technology:
The Second Quantum Revolution, quant-ph/0206091.
[8] C. H. Bennett and D. P. DiVincenzo, Nature 404, 247
(2000).
[9] M. Hennrich, T. Legero, A. Kuhn, and G. Rempe, Phys.
Rev. Lett. 85, 4872 (2000); A. Kuhn, M. H ennrich, and
G. Rempe, Phys. Rev. Lett. 89, 067901 (2002).
[10] J. McKeever, A. Boca, A . D. Boozer, J. R. Buck, and H.
J. Kimble, Nature 425, 268 (2003).
[11] J. A. Sauer, K. M. Fortier, M. S. Chang, C. D. Hamley,
and M. S. Chapman, Cavity QED with optically trans-
ported atoms, quant-ph/0309052.
[12] G. R. Guth¨ohrlein, M. Keller, K. Hayasaka, W. Lange,
and H. Walther, Nature 414, 49 (2001).
[13] C. Sch¨on and J. I. Cirac, Phys. Rev. A 67, 043813 (2003).
[14] T. Fischer, P. Maunz, P. W . H. Pinkse, T. Puppe, and
G. Rempe, Phys. Rev. Lett. 88, 163002 (2002).
[15] P. Horak, B. G. Klappauf, A. Haase, R. Folman, J.
Schmiedmayer, P. Domokos, and E. A. Hinds, Phys. Rev.
A 67, 043806 (2003).
[16] T. Pellizzari, S. A. Gardiner, J. I. Cirac, and P. Zoller,
Phys. Rev. Lett. 75, 3788 (1995).
[17] P. Domokos, J. M. Raimond, M. Brune, and S. Haroche,
Phys. Rev. A 52, 3554 (1995).
[18] A. Beige, D. Braun, B. Tregenna, and P. L. Knight, Phys.
Rev. Lett. 85, 1762 (2000); B. Tregenna, A. Beige, and
P. L. Knight, Phys. Rev. A 65, 032305 (2002).
[19] S.-B. Zheng and G.-C. Guo, Phys. Rev. Lett. 85, 2392
(2000).
[20] J. Pachos and H. Walther, Phys. Rev. Lett . 89, 187903
(2002).
[21] E. Jan´e, M. B. Plenio, and D. Jonathan, Phys. Rev. A
65, 050302 (2002).
[22] A. Recati, T. Calarco, P. Zanardi, J. I. Cirac, and P.
Zoller, Phys. Rev. A 66, 032309 (2002).
[23] L. You, X. X. Yi, and X. H. Su, Phys. Rev. A 67, 032308
(2003); X. X. Yi, X. H. Su, and L. You, Phys. Rev. Lett.
90, 097902 (2003).
[24] A. S. Sørensen and K. Mølmer, Phys. Rev. Lett. 91,
097905 (2003).
[25] M. V. Berry, Proc. R. Soc. London S er. A 392, 45 (1984).
[26] G. Palma, K.-A. Suominen, and A. Ekert, Proc. Roy.
Soc. London Ser. A 452, 567 (1996).
[27] P. Zanardi and M. Rasetti, Phys. Rev. Lett. 79, 3306
(1997).
[28] D. A . Lidar, I. L. Chuang, and K. B. Whaley, Phys. Rev.
Lett. 81, 2594 (1998).
[29] A. Beige, D. Braun, and P. L. Knight, New J. Phys. 2,
22 (2000).
[30] J. Oreg, F. T. Hioe, and J. H. Eberly, Phys. Rev. A 29,
690 (1984).
[31] K. Bergmann, H. Theuer, and B. W. Shore, Rev. Mod.
Phys. 70, 1003 (1998); N. V. Vitanov, T. Halfmann, B.
W. Shore, and K. Bergmann, Annu. Rev. Phys. Chem.
52, 763 (2001).
[32] C. Marr, A. Beige, and G. R empe, Phys. Rev. A 68,
033817 (2003).
[33] For a detailed description of quantum computing using
dissipation see A. Beige, Phys. Rev. A (in press), Section
II; quant-ph/0304168.
[34] G. C. Hegerfeldt and T. S. Wilser, in Classical and Quan-
tum Systems, Proceedings of the Second International
Wigner Symposium, July 1991, ed ited by H. D. Doebn er,
W. Scherer, and F. Schroeck (World Scientific, S ingapore,
1992), p. 104; G. C. Hegerfeldt and D. G. Sondermann,
Quant. Semiclass. Op t. 8, 121 (1996).
[35] J. Dalibard, Y. Castin, and K. Mølmer, Phys. Rev. Lett.
68, 580 (1992).
[36] H. Carmichael, An Open Systems Approach to Quan-
tum Optics, Lecture Notes in Physics, Vol. 18 (Springer,
Berlin, 1993).
[37] E. Knill, R . Laflamme, and G. J. Milburn, Nature 409,
46 (2001).
[38] P. Zanardi and M. R asetti, Phys. Lett. A 264, 94
9
(1999); J. Pachos, “Quantum Computation by Geomet-
rical Means”, published in the AMS Contemporary Math
Series volume entitled “Quantum Computation & Quan-
tum In formation Science”.
[39] D. Bouwmeester, G. P. Karman, N. H. Dekker, C. A.
Schramm, and J. P. Woerdman, J. Mod. Opt. 43, 2087
(1996).
[40] G. G. Lapaire, P. Kok, J. P. Dowling, and J. E. Sipe,
Conditional linear-optical measurement schemes gener-
ate effective photon nonlinearities, quant-ph/0305152.
[41] D. Ellinas and J. Pachos, Phys. Rev. A 64, 022310 (2001).
[42] A. Nazir, T. P. Spiller, and W. J. Munro, Phys. Rev. A
65, 042303 (2002).
[43] A. Carollo, I. Fuentes-Guridi, M. F. Santos, and V. Ve-
dral, Phys. Rev. Lett. 90, 160402 (2003).
10
... It is robust against errors in the laser pulse parameters and it is highly immune to spontaneous decay and dephasing of the intermediate state [7,8]. There are several proposed ways to use STIRAP for quantum gates and entanglement in different types of quantum systems, including Rydberg atoms [9][10][11][12][13][14][15][16][17][18][19][20][21][22], spin systems [23], cold ions [24] and qubits interacting via a cavity [25][26][27][28][29]. There are also experimental demonstrations of quantum gates and entanglement in cold ions [30,31] and in cold atoms [32,33] using STIRAP or similar adiabatic transfer methods. ...
Article
Full-text available
This paper investigates the generation of quantum entanglement by means of conditional stimulated Raman adiabatic passage (STIRAP) based on Rydberg blockade. The paper compares the entanglement fidelities in three-level and four-level schemes and analyzes the adiabatic conditions in both cases. In particular, Green–Horne–Zeilinger states can be deterministically generated in an atomic ensemble interacting with a single control atom.
... Due to such advantages, STIRAP and its many variants have quickly evolved into general methods of quantum manipulation beyond the original usage for state transfer, and found wide applications in many subfields of physics [5], chemistry [6], and engineering [5,7]. One particular field is quantum information processing, where STIRAP has been proposed and demonstrated for implementing quantum gates [8][9][10][11][12][13][14][15][16], state preparation and transfer [17][18][19][20][21], quantum computation in extended Hilbert space [22,23], quantum memory [24,25], and so on. However, the demand of high precision and fidelity in this field, as well as in other fields where STIRAP may find potential applications [5], poses great challenges to the original format of STIRAP. ...
Article
Full-text available
Stimulated Raman adiabatic passage (STIRAP) is now a standard technique of coherent state-to-state manip-ulation for many applications in physics, chemistry, and beyond. The adiabatic evolution of the state involvedin STIRAP, called adiabatic passage, guarantees its robustness against control errors. However, the applicabilityof STIRAP is limited by the problems of low efficiency and decoherence. Here we propose, and experimen-tally demonstrate, an alternative approach, termed stimulated Raman “user-defined” passage (STIRUP), wherea customizable state is employed for constructing desired evolutions to replace the adiabatic passage in STI-RAP. The user-defined passages can be flexibly designed for optimizing different objectives for different tasks,such as minimizing leakage errors. To experimentally benchmark its performance, we apply STIRUP to thetask of coherent state transfer in a superconducting Xmon qutrit. We verified that STIRUP can complete thetransfer at least four times faster than STIRAP with enhanced robustness, achieving a fidelity of 99.5%, whichis highest among all recent experiments based on STIRAP and its variants. In practice, STIRUP differs fromSTIRAP only in the design of driving pulses; therefore, most existing applications of STIRAP can be readilyimplemented using STIRUP with a potential improvement in performance .
... Due to such advantages, STIRAP and its many variants have quickly evolved into general methods of quantum manipulation beyond the original usage for state transfer, and found wide applications in many subfields of physics [5], chemistry [6], and engineering [5,7]. One particular field is quantum information processing, where STIRAP has been proposed and demonstrated for implementing quantum gates [8][9][10][11][12][13][14][15][16], state preparation and transfer [17][18][19][20][21], quantum computation in extended Hilbert space [22,23], quantum memory [24,25], and so on. However, the demand of high precision and fidelity in this field, as well as in other fields where STIRAP may find potential applications [5], poses great challenges to the original format of STIRAP. ...
Preprint
Stimulated Raman adiabatic passage (STIRAP) is now a standard technique of coherent state-to-state manipulation for many applications in physics, chemistry, and beyond. The adiabatic evolution of the state involved in STIRAP, called adiabatic passage, guarantees its robustness against control errors. However, the applicability of STIRAP is limited by the problems of low efficiency and decoherence. Here we propose, and experimentally demonstrate, an alternative approach, termed stimulated Raman "user-defined" passage (STIRUP), where a customizable state is employed for constructing desired evolutions to replace the adiabatic passage in STIRAP. The user-defined passages can be flexibly designed for optimizing different objectives for different tasks, such as minimizing leakage errors. To experimentally benchmark its performance, we apply STIRUP to the task of coherent state transfer in a superconducting Xmon qutrit. We verified that STIRUP can complete the transfer at least four times faster than STIRAP with enhanced robustness, achieving a fidelity of 99.5%, which is highest among all recent experiments based on STIRAP and its variants. In practice, STIRUP differs from STIRAP only in the design of driving pulses; therefore, most existing applications of STIRAP can be readily implemented using STIRUP with a potential improvement in performance.
... Cavity quantum electrodynamics (QED) as a platform for processing quantum information has attracted persistent interest. Based on cavity QED several theoretical proposals have been presented for implementation quantum phase gate [17][18][19][20]. In particular, Duan and Kimble [21] proposed an interesting scheme to carry out controlled-phase-flip (CPF) gate between the single-photon pulses through the cavity input-output process. ...
Article
Full-text available
Multiqubit quantum controlled-phase-flip (CPF) gate between atomic qubits is desirable for scalable and distributed quantum computation. Here, we propose a scheme to realize a multiqubit quantum CPF gate between different atoms, which are trapped in separate cavities coupled by short optical fiber. After a single-photon pulse reflected by the cavity-atoms system, a multiqubit CPF gate can be implemented by only one step.
... By properly encoding the physical qubits, decoherence which comes from collective dephasing can be canceled. In order to deal with both control errors and collective dephasing noises, the schemes of nonadiabatic GQC-DFSs have been proposed [24][25][26][27]36,[43][44][45][46][47][48]. However, these schemes are still sensitive to the systematic errors due to the nonadiabatic feature. ...
Article
Full-text available
We propose a practical scheme to implement universal superadiabatic geometric quantum gates in decoherence-free subspaces in the trapped-ions system. The logical qubit is only encoded by two neighboring physical qubits, which is the minimal resource for the decoherence-free subspace encoding. Different from the nonadiabatic control in decoherence-free subspace (Liang et al. in Phys Rev A 89:062312, 2014), a new Hamiltonian to implement universal effective interaction between logical qubits is proposed in the scheme. The proposed gates are numerically demonstrated to be robust against both systematic errors and collective dephasing noises, which combine the advantages of superadiabatic geometric quantum control and decoherence-free subspace. Since the Hamiltonian we use relies solely on two-body interactions, our scheme would be promising to be realized experimentally in trapped-ions systems.
... The method avoids both spontaneous emission, because of STIRAP, and cavity loss because no photon is present in the cavity at any time. A modified version of this method (Pachos and Beige, 2004) creates a two-qubit phase gate, with either dynamical or geometric phase, by using a common laser addressing of the two qubits in a single step. Goto and Ichimura (2004) proposed STIRAPinspired implementations of one-, two-and three-qubit phase gates of atoms in a single-mode optical cavity. ...
Article
Full-text available
Adiabatic techniques are known to allow for engineering quantum states with high fidelity. This requirement is currently of large interest, as applications in quantum information require the preparation and manipulation of quantum states with minimal errors. Here we review recent progress on developing techniques for the preparation of spatial states through adiabatic passage, particularly focusing on three state systems. These techniques can be applied to matter waves in external potentials, such as cold atoms or electrons, and to classical waves in waveguides, such as light or sound.
... The method avoids both spontaneous emission, because of STIRAP, and cavity loss because no photon is present in the cavity at any time. A modified version of this method (Pachos and Beige, 2004) creates a two-qubit phase gate, with either dynamical or geometric phase, by using a common laser addressing of the two qubits in a single step. Goto and Ichimura (2004) proposed STIRAPinspired implementations of one-, two-and three-qubit phase gates of atoms in a single-mode optical cavity. ...
Article
Full-text available
The technique of stimulated Raman adiabatic passage (STIRAP), which allows efficient and selective population transfer between quantum states without suffering loss due to spontaneous emission, was introduced in 1990 (Gaubatz \emph{et al.}, J. Chem. Phys. \textbf{92}, 5363, 1990). Since then STIRAP has emerged as an enabling methodology with widespread successful applications in many fields of physics, chemistry and beyond. This article reviews the many applications of STIRAP emphasizing the developments since 2000, the time when the last major review on the topic was written (Vitanov \emph{et al.}, Adv. At. Mol. Opt. Phys. \textbf{46}, 55, 2001). A brief introduction into the theory of STIRAP and the early applications for population transfer within three-level systems is followed by the discussion of several extensions to multi-level systems, including multistate chains and tripod systems. The main emphasis is on the wide range of applications in atomic and molecular physics (including atom optics, cavity quantum electrodynamics, formation of ultracold molecules, precision experiments, etc.), quantum information (including single- and two-qubit gates, entangled-state preparation, etc.), solid-state physics (including processes in doped crystals, nitrogen-vacancy centers, superconducting circuits, etc.), and even some applications in classical physics (including waveguide optics, frequency conversion, polarization optics, etc.). Promising new prospects for STIRAP are also presented (including processes in optomechanics, detection of parity violation in molecules, spectroscopy of core-nonpenetrating Rydberg states, and population transfer with X-ray pulses).
... The method avoids both spontaneous emission, because of STIRAP, and cavity loss because no photon is present in the cavity at any time. A modified version of this method (Pachos and Beige, 2004) creates a two-qubit phase gate, with either dynamical or geometric phase, by using a common laser addressing of the two qubits in a single step. Goto and Ichimura (2004) proposed STIRAPinspired implementations of one-, two-and three-qubit phase gates of atoms in a single-mode optical cavity. ...
Article
Full-text available
The technique of stimulated Raman adiabatic passage (STIRAP), which allows efficient and selective population transfer between quantum states without suffering loss due to spontaneous emission, was introduced in 1990 (Gaubatz et al., J. Chem. Phys. 92, 5363, 1990). Since then STIRAP has emerged as an enabling methodology with widespread successful applications in many fields of physics, chemistry and beyond. This article reviews the many applications of STIRAP emphasizing the developments since 2000, the time when the last major review on the topic was written (Vitanov et al., Adv. At. Mol. Opt. Phys. 46, 55, 2001). A brief introduction into the theory of STIRAP and the early applications for population transfer within three-level systems is followed by the discussion of several extensions to multi-level systems, including multistate chains and tripod systems. The main emphasis is on the wide range of applications in atomic and molecular physics (including atom optics, cavity quantum electrodynamics, formation of ultracold molecules, precision experiments, etc.), quantum information (including single- and two-qubit gates, entangled-state preparation, etc.), solid-state physics (including processes in doped crystals, nitrogen-vacancy centers, superconducting circuits, etc.), and even some applications in classical physics (including waveguide optics, frequency conversion, polarization optics, etc.). Promising new prospects for STIRAP are also presented (including processes in optomechanics, detection of parity violation in molecules, spectroscopy of core-nonpenetrating Rydberg states, and population transfer with X-ray pulses).
Article
Adiabatic quantum-control schemes are widely used in the areas of quantum-information processing and quantum sensing. The very requirement of adiabaticity in these schemes often leads to problems of low efficiency and being sensitive to decoherence. To address such issues, various methods have been developed for accelerating adiabatic processes, but they often need to be designed on an adhoc basis. Here we propose and experimentally demonstrate an inverse-engineering approach, where a parameterized state of the Schrödinger equation is employed for constructing desired evolutions. Within a single and simple frame, the user-defined passages can be flexibly customized for different objectives and systems. To experimentally benchmark its performance, we implement this approach in a task of coherent state transfer in a superconducting Xmon device. We find that our method completed the transfer with the fastest speed and highest efficiency (>99.5%) among all recent experiments performed in similar configurations.
Article
Universal computation of a quantum system consisting of superpositions of well-separated coherent states of multiple harmonic oscillators can be achieved by three families of adiabatic holonomic gates. The first gate consists of moving a coherent state around a closed path in phase space, resulting in a relative Berry phase between that state and the other states. The second gate consists of “colliding” two coherent states of the same oscillator, resulting in coherent population transfer between them. The third gate is an effective controlled-phase gate on coherent states of two different oscillators. Such gates should be realizable via reservoir engineering of systems that support tunable nonlinearities, such as trapped ions and circuit QED.
Article
Full-text available
The effects of certain forms of decoherence applied to adiabatic and nonadiabatic geometric phase quantum gates were studied. Effects of isotropic noise were set by the gate length and level of decoherence. Results indicated that the anisotropic noise-sensitive gates coupled with the ability to reorienate the external static magnetic field, could be applied in mapping out the forms of decoherence acting on a qubit. The single-qubit decoherence calibrations could be used to predict the expected level of entanglement in two-qubit gates prior to experiment.
Article
Full-text available
We investigate the optical detection of single atoms held in a microscopic atom trap close to a surface. Laser light is guided by optical fibers or optical microstructures via the atom to a photodetector. Our results suggest that with present-day technology microcavities can be built around the atom with sufficiently high finesse to permit unambiguous detection of a single atom in the trap with 10μs of integration. We compare resonant and nonresonant detection schemes and discuss the requirements for detecting an atom without causing it to undergo spontaneous emission.
Article
Full-text available
By manipulating the discrete optical levels inside an optical resonator, we obtain a classical realization of a twisted Landau-Zener model. We experimentally demonstrate the geometric amplitude factor in the transition amplitude that arises for this model. We consider in particular the region of parameter space addressed in the original study of the geometric amplitude factor by M. V. Berry [7].
Article
It is 200 years since Thomas Young performed his famous double-slit experiment but the interference of waves that weave rich tapestries in space and space-time. continues to provide deep insights into geometrical optics and semi-classical limits.
Article
We propose and analyze an experiment designed to probe the extent to which information accessible to an observer and the "eraser" of this information affects measured results. The proposed experiment could also be operated in a "delayed-choice" mode.
Article
A quantal system in an eigenstate, slowly transported round a circuit C by varying parameters R in its Hamiltonian hat{H}(R), will acquire a geometrical phase factor exp {igamma(C)} in addition to the familiar dynamical phase factor. An explicit general formula for gamma(C) is derived in terms of the spectrum and eigenstates of hat{H}(R) over a surface spanning C. If C lies near a degeneracy of hat{H}, gamma(C) takes a simple form which includes as a special case the sign change of eigenfunctions of real symmetric matrices round a degeneracy. As an illustration gamma(C) is calculated for spinning particles in slowly-changing magnetic fields; although the sign reversal of spinors on rotation is a special case, the effect is predicted to occur for bosons as well as fermions, and a method for observing it is proposed. It is shown that the Aharonov-Bohm effect can be interpreted as a geometrical phase factor.
Article
The authors discuss the technique of stimulated Raman adiabatic passage (STIRAP), a method of using partially overlapping pulses (from pump and Stokes lasers) to produce complete population transfer between two quantum states of an atom or molecule. The procedure relies on the initial creation of a coherence (a population-trapping state) with subsequent adiabatic evolution. The authors present the basic theory, with some extensions, and then describe examples of experimental utilization. They note some applications of the technique not only to preparation of selected states for reaction studies, but also to quantum optics and atom optics.
Article
The problem of achieving population inversion adiabatically in an N-level system using one or more laser fields whose detunings and/or amplitudes are continuously varied is studied analytically and numerically. The SU(N) coherence vector picture is shown to suggest unexpected inversion procedures and also to give a generalized interpretation of adiabatic following. It is shown that the (N2-1)-dimensional SU(N) space contains an (N-1)--dimensional steady-state subspace Γ(t) whose orthonormal basis vectors Γ⃗1,…, Γ⃗N-1 are given explicitly in terms of the Hamiltonian matrix elements. The motion of the system can be interpreted as a "generalized precession" of S⃗ about Γ. Multilevel adiabatic following occurs when the angle χ(t) between the coherence vector S⃗ and its projection onto Γ is very small. The multiple dimension of Γ is shown to provide a variety of paths for adiabatic inversion. The adiabatic solution is obtained by solving N-1 simple equations for the directional cosines of S⃗ on Γ⃗i. The adiabatic solution and time scale and the state taken up by the atomic variable are discussed analytically and numerically for a three-level system.
Article
The aim of this work is to find ways to trap an atom in a cavity. In contrast to other approaches, we propose a method where the cavity is basically in the vacuum state and the atom is in the ground state. The idea is to induce a spatial dependent ac-Stark shift by irradiating the atom with a weak laser field, so that the atom experiences a trapping force. The main feature of our setup is that dissipation can be strongly suppressed. We estimate the lifetime of the atom as well as the trapping potential parameters, and compare our estimations with numerical simulations.