Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[J. Res. Natl. Inst. Stand. Technol. 103, 259 (1998)]
Experimental Issues in Coherent Quantum-State
Manipulation of Trapped Atomic Ions
Volume 103 Number 3 May–June 1998
D. J. Wineland, C. Monroe,
W. M. Itano, D. Leibfried
1
,
B. E. King, and D. M. Meekhof
National Institute of Standards and
Technology,
Boulder, CO 80303
Methods for, and limitations to, the genera-
tion of entangled states of trapped atomic
ions are examined. As much as possible,
state manipulations are described in terms
of quantum logic operations since the con-
ditional dynamics implicit in quantum logic
is central to the creation of entanglement.
Keeping with current interest, some experi-
mental issues in the proposal for trapped-
ion quantum computation by J. I. Cirac and
P. Zoller (University of Innsbruck) are dis-
cussed. Several possible decoherence mech-
anisms are examined and what may be the
more important of these are identified.
Some potential applications for entangled
states of trapped-ions which lie outside the
immediate realm of quantum computation
are also discussed.
Key words: coherent control; entangled
states; laser cooling and trapping; quantum
computation; quantum state engineering;
trapped ions.
Accepted: February 4, 1998
Available online: http://www.nist.gov/jres
Contents
1. Introduction ............................... 260
2. Trapped Atomic Ions........................ 261
2.1. Ions Confined in Paul Traps ............. 261
2.2. Ion Motional and Internal Quantum States. . 263
2.2.1. Detection of Internal States ........ 264
2.3. Interaction With Additional Applied
Electromagnetic Fields.................. 265
2.3.1. Single Ion, Single Applied Field,
Single Mode of Motion ........... 265
2.3.2. State Dynamics Including Multiple
Modes of Motion ................ 268
2.3.3. Stimulated-Raman Transitions ...... 269
3. Quantum-State Manipulation.................. 270
3.1. Laser Cooling to the Ground State of
Motion .............................. 270
3.2. Generation of Nonclassical States of
Motion of a Single Ion ................. 272
3.2.1. Population Analysis of Motional
States.......................... 272
3.2.2. Fock State ...................... 272
3.2.3. Coherent States .................. 273
3.2.4. Other Nonclassical States.......... 274
3.3. Quantum Logic ....................... 274
1
Present address: Universita¨t Innsbruck, Institut fu¨r Experimental
Physik, Austria.
3.4. Entangled States for Spectroscopy ........ 278
4. Decoherence............................... 279
4.1. Motional Decoherence.................. 279
4.1.1. Phase Decoherence Caused by
Unstable Trap Parameters......... 280
4.1.2. Radiative Decoherence ........... 280
4.1.3. Radiative Damping/Heating ....... 280
4.1.4. Injected Noise .................. 282
4.1.5. Motional Excitation From Trap
rfFields....................... 283
4.1.6. Fluctuating Patch Fields .......... 285
4.1.7. Field Emission.................. 286
4.1.8. Mode Cross Coupling From Static
Electric Field Imperfections ....... 286
4.1.9. Collisions With Background Gas . . . 287
4.1.10. Experimental Studies of Heating . . . 289
4.1.11. Experimental Studies of Motional
Decoherence ................... 289
4.2. Internal State Decoherence .............. 290
4.2.1. Radiative Decay ................ 291
4.2.2. Magnetic Field Fluctuations ....... 291
4.2.3. Electric Field Fluctuations ........ 292
4.3 Logic Operation Fidelity and Rotation
Angle Errors.......................... 293
4.3.1 Accumulated Errors ............. 294
4.3.2. Pulse Area and Phase Fluctuations . 296
4.4. Sources of Induced Decoherence ......... 297
259
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.4.1. Applied Field Amplitude and
Timing Fluctuations............... 297
4.4.2. Characterization of Amplitude
and Timing Fluctuations ........... 297
4.4.3. Applied Field Frequency and Phase
Fluctuations ..................... 299
4.4.4. Individual Ion Addressing and
Applied Field Position Sensitivity.... 301
4.4.5. Effects of Ion Motion (Debye-Waller
Factors) ........................ 303
4.4.6. Coupling to Spectator Levels ....... 305
4.4.6.1. Polarization Discrimination
of Internal States .......... 305
4.4.6.2. Spectral Discrimination
of States ................. 306
4.4.6.3. Tailoring of Laser Fields .... 307
4.4.6.4. Spontaneous Emission ...... 308
4.4.7. Mode Cross Coupling During Logic
Operations ...................... 309
5. Variations ................................. 310
5.1. Few Ion Accumulators ................. 310
5.2. Multiplexing With Internal States ......... 311
5.3. High-Z Hyperfine Transitions ............ 311
6. Other Applications.......................... 312
6.1. Quantum Correlations .................. 312
6.2. Simulations........................... 313
6.2.1. Mach-zehnder Boson Interferometer
With Entangled States ............. 314
6.2.2. Squeezed-Spin States.............. 315
6.3. Mass Spectrometry and NMR at the Single
Quantum Level........................ 316
6.4. Quantum State Manipulation of Mesoscopic
Mechanical Resonators ................. 319
7. Summary/Conclusions ....................... 320
8. Appendix A. Entangled States and Atomic
Clocks.................................... 321
9. Appendix B. Master Equation for the Density
Matrix of a Radiatively Damped Harmonic
Oscillator ................................. 322
10. References
............................. 323
1. Introduction
A number of recent theoretical and experimental
papers have investigated the ability to coherently
control or “engineer” atomic, molecular, and optical
quantum states. This theme is manifested in topics such
as atom interferometry, atom optics, the atom laser,
Bose-Einstein condensation, cavity QED, electromag-
netically induced transparency, lasing without inver-
sion, quantum computation, quantum cryptography,
quantum-state engineering, squeezed states, and
wavepacket dynamics. In this paper, we investigate a
subset of these topics which involve the coherent
manipulation of quantum states of trapped atomic ions.
The focus will be on a proposal to implement quantum
logic and quantum computation using trapped ions [1].
However, we will also consider related work on the
generation of nonclassical states of motion and entan-
gled states of trapped ions [2–39]. Many of these ideas
have been summarized in a recent review [40].
Coherent control of spins and internal atomic states
has a long history in NMR and rf /laser spectroscopy.
For example, the ability to realize coherent
p
pulses”
or
p
/2 pulses” on two-level systems has been routine
for decades. In much of what is discussed in this paper,
we will consider entangling operations, that is, unitary
operations which create entangled states between two or
more separate quantum systems. In particular, we will
be interested in situations where the interaction between
quantum systems can be selectively turned on and off.
For brevity, we will limit discussion to these types of
operations in experiments which involve trapped
atomic ions; however, many of the discussions, in
particular those concerning single trapped ions, will
also apply to trapped neutral atom experiments where
the atoms can be treated as independent. The aspect of
entangling operations is shared by atom optics and atom
interferometry [41, 42] and, as described below, there
are close parallels between the ion trap experiments and
those of cavity QED [43].
Earlier experiments on trapped ions, where the zero-
point of motion was closely approached through laser
cooling, already showed the effects of nonclassical
motion in the absorption spectrum [44–46]. These same
effects can be used to characterize the average energy of
the ion. More recent experiments report the generation
of Fock, squeezed, coherent [21], and Schro¨dinger cat
[47] states. These states appear to be of fundamental
physical interest and possibly of use for sensitive detec-
tion of small forces [26, 48]. For comparison, experi-
ments which detect quantized atomic motion in optical
lattices are reviewed by Jessen and Deutsch [49]. Also,
through the mechanism of Bose-Einstein condensation,
which has recently been observed in neutral atomic
vapors [50–53], a macroscopic occupation of a single
motional state (the ground state of motion) is achieved.
Simple quantum logic experiments have been carried
out with single trapped ions [17]; the emphasis of future
work will be to implement quantum logic on many ions
[1]. The attendant ability to create correlated, or entan-
gled, states of atomic particles appears to be interesting
from the standpoint of quantum measurement [54] and,
for example, for improved signal-to-noise ratio in
spectroscopy of trapped ions (Sec. 3.4).
Therefore, we will be particularly interested in study-
ing the practical limits of applying coherent control
methods to trapped ions for (1) the generation and
analysis of nonclassical states of motion, (2) the imple-
mentation of quantum logic and computation, and (3)
the generation of entangled states which can improve
260
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
signal-to-noise ratio in spectroscopy. We will briefly
describe the experimental results in these three areas,
but the main purpose of the paper will be to anticipate
and characterize decohering mechanisms which limit
the ability to produce the desired final quantum states in
current and future experiments. This is a particularly
important issue for quantum computation where many
ions (thousands) and coherent operations (billions) may
be required in order for quantum computation to be
generally useful. Here, we generalize the meaning of
decoherence to include any effect which limits the
purity of the desired final states. A fundamental source
of decoherence will be the coupling of the ion’s motion
and internal states to the environment. Also important is
induced decoherence caused by, for example, technical
fluctuations in the applied fields used to implement the
operations. This division between types of decoherence
is arbitrary since both effects can be regarded as
coupling to the environment; however, the division will
provide a useful framework for discussion. As a unify-
ing theme for the paper, we will find it useful to regard,
as much as possible, the quantum manipulations we
discuss in terms of quantum logic. Of course, the subject
of decoherence is much broader than the specific
context discussed here; the reader is referred to more
general discussions such as the papers by Zurek
[55,56,57].
The paper is organized as follows. In the next section,
we briefly discuss ion trapping. In Sec. 3, we consider
in somewhat more detail the three areas of application
enumerated in the previous paragraph. Since cooling of
the ions to their ground state of motion is a prerequisite
to the main applications discussed in the paper, we out-
line methods to accomplish this in the beginning of Sec.
3. Section 4 is the heart of the paper; here, we attempt
to identify the most important sources of decoherence.
Section 5 briefly discusses some variations on proposed
methods for realizing quantum logic in trapped ions.
Section 6 suggests some additional applications of the
ideas discussed in the paper and Sec. 7 provides a brief
summary.
Such a treatment seems warranted in that several lab-
oratories are investigating the use of trapped ions for
quantum logic and related topics; the authors are aware
of related experiments being pursued at IBM, Almaden;
Innsbruck University; Los Alamos National Labora-
tory; Max Planck Institute, Garching; NIST, Boulder;
and Oxford University. This analysis in this paper
necessarily overlaps, but is also intended to complement,
other investigations [58–66] and will, by no means, be
the end of the story. We hope however, that this paper
will stimulate others to do more complete treatments
and consider effects that we have neglected.
2. Trapped Atomic Ions
2.1 Ions Confined in Paul Traps
Due to their net charge, atomic ions can be confined
by particular arrangements of electromagnetic fields.
For studies of ions at low energy, two types of trap are
typically used—the Penning trap, which uses a combi-
nation of static electric and magnetic fields, and the Paul
or rf trap which confines ions primarily through pon-
deromotive forces generated by inhomogeneous oscil-
lating fields. The operation of these traps is discussed in
various reviews (see for example, Refs. [67]–[70]), and
in a recent book by Ghosh [71]. For brevity, we discuss
one trap configuration, the linear Paul trap, which may
be particularly useful in the context of this paper. This
choice however, does not rule out the use of other types
of ion traps for the experiments discussed here.
In Fig. 1 we show a schematic diagram of a linear
Paul trap. This trap is based on the one described by
Raizen et al. [72] which is derived from the original
design of Drees and Paul [73]. It is basically a quadru-
pole mass filter which is plugged at the ends with static
electric potentials. A potential V
0
cos
V
T
t + U
r
is applied
between diagonally opposite rods, which are fixed in a
quadrupolar configuration, as indicated in Fig. 1. We
assume that the rod segments along the z direction are
coupled together with capacitors (not shown) so that the
rf potential is constant as a function of z. Near the axis
of the trap this creates a potential of the form
F
.
(V
0
cos
V
T
t + U
r
)
2
S
1+
x
2
y
2
R
2
D
, (1)
where R is equal to the distance from the axis to the
surface of the electrode. (Unless the rods conform to
equipotentials of Eq. (1), this equation must be multi-
plied by a constant factor on the order of 1; see for
example, Ref. [72].) This gives rise to (harmonic)
ponderomotive potentials in the x and y directions. To
provide confinement along the z direction, static poten-
tials U
0
are applied to the end segments of the rods as
indicated. Near the center of the trap, this provides a
static harmonic well in the z direction
F
s
= kU
0
F
z
2
1
2
(x
2
+ y
2
)
G
=
m
2q
v
2
z
F
z
2
1
2
(x
2
+ y
2
)
G
,
(2)
where k is a geometric factor, m and q are the ion mass
and charge, and
v
z
=(2kqU
0
/m)
1/2
is the oscillation
frequency for a single ion or the center-of-mass (COM)
261
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
sections, a
i
(or a) will represent the harmonic oscillator
lowering operator and q
i
will represent the normal mode
coordinate for the ith mode). The solution of Eqs. (3) to
first order in a
i
and second order in q
i
is given by
u
i
(t)=A
i
S
cos(
v
i
t +
f
i
)
F
1+
q
i
2
cos(
V
T
t)
+
q
2
i
32
cos(2
V
T
t)
G
+
b
i
q
i
2
sin(
v
i
t +
f
i
) sin(
V
T
t)
D
, (4)
where u
i
= x or y, A
i
depends on initial conditions, and
v
i
=
b
i
V
T
2
,
b
i
. [a
i
+ q
2
i
/2]
1/2
. (5)
The large amplitude oscillation at frequency
v
i
is typi-
cally called the “secular” motion. When a
i
<< q
2
i
<< 1
and U
r
. 0, if we neglect the micromotion (the terms
which oscillate at
V
T
and 2
V
T
), the ion behaves as if it
were confined in a harmonic pseudopotential
F
p
in the
radial direction given by
q
F
p
=
1
2
m
v
2
r
(x
2
+ y
2
) , (6)
where
v
r
. qV
0
/(2
1/2
V
T
mR
2
)=q
x
V
T
/(2Ï2) is the radial
secular frequency
v
r
. For most of the discussions in this
paper, we will assume U
r
= 0; however it may be useful
in some cases to make U
r
Þ 0 to break the degeneracy
of the x and y frequencies. Figure 1 also shows an image
of a “string” of
199
Hg
+
ions which are confined near the
z axis of the trap described in Ref. [74]. This was
achieved by making
v
r
>>
v
z
, thereby forcing the ions
to the axis of the trap. The spacings between individual
ions in this string are governed by a balance of the force
along the z direction due to
F
s
and the mutual Coulomb
repulsion of the ions. Example parameters are given in
the figure caption.
When this kind of trap is installed in a high-vacuum
apparatus, ions can be confined for days with minimal
perturbations to their internal structure. Collisions with
background gas can be neglected (Sec. 4.1.9). Even
though the ions interact strongly through their mutual
Coulomb interaction, the fact that the ions are localized
necessarily means that the time-averaged value of the
electric field they experience is zero; therefore electric
field perturbations are small (Sec. 4.2.3). Magnetic field
perturbations to internal structure are important; how-
ever, the coherence time for superposition states of two
internal levels can be very long by operating at fields
Fig. 1. The upper part of the figure shows a schematic diagram of
the electrode configuration for a linear Paul-rf trap (rod spacing .
1 mm). The lower part of the figure shows an image of a string of
199
Hg
+
ions, illuminated with 194 nm radiation, taken with a uv-sensi-
tive, photon counting imaging tube [74]. The spacing between adja-
cent ions is approximately 10 mm. The “gaps” in the string are occu-
pied by impurity ions, most likely other isotopes of Hg
+
, which do not
fluoresce because the frequencies of their resonant transitions do not
coincide with those of the 194 nm
2
S
1/2
2
P
1/2
transition of
199
Hg
+
.
oscillation frequency for a collection of identical ions
along the z direction. Equations (1) and (2) represent the
lowest order terms in the expansion of the potentials for
the electrode configuration of Fig. 1. When the size of
the ion sample or amplitude of ion motion is
comparable to the spacing between electrodes or the
spacing between rod segments, higher order terms in
F
and
F
s
become important. However for small oscilla-
tions of the COM mode, which is relevant here, the
harmonic approximation will be valid. In the x and y
directions, the action of the potentials of Eqs. (1) and (2)
gives the (classical) equations of motion described by
the Mathieu equation
d
2
x
d
z
2
+
F
a
x
+2q
x
cos(2
z
)
G
x =0
d
2
y
d
z
2
+
F
a
y
+2q
y
cos(2
z
)
G
y = 0, (3)
where
z
V
T
t/2, a
x
=(4q/m
V
2
T
)(U
r
/R
2
kU
0
/z
2
0
),
a
y
=–(4q/m
V
2
T
)(U
r
/R
2
+ kU
0
/z
2
0
), q
x
=–q
y
=2
q
V
0
/
(
V
2
T
mR
2
). The Mathieu equation can be solved in
general using Floquet solutions. Typically, we will have
a
i
< q
2
i
<< 1, i [ {x, y}. (Keeping with the usual
notation in the ion-trap literature, in this section, the
symbols a
i
and q
i
are defined as above. In all other
262
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
where the energy separation between levels is at an
extremum with respect to field. For example, in a
9
Be
+
(Penning trap) experiment operating in a field of 0.82 T,
a coherence time between hyperfine levels exceeding 10
min was observed [75, 76]. As described below, we will
be interested in coherently exciting the quantized modes
of the ions’ motion in the trap. Here, not surprisingly,
the coupling to the environment is relatively strong be-
cause of the ions’ charge. One measure of the decoher-
ence rate is obtained from the linewidth of observed
motional resonances of the ions; this gives an indication
of dephasing times. For example, the linewidths of
cyclotron resonance excitation in high resolution mass
spectroscopy in Penning traps [77–79] indicate that
these coherence times can be at least as long as several
tens of seconds. Decoherence can also occur from tran-
sitions between the ions’ quantized oscillator levels.
Transition times out of the zero-point motional energy
level have been measured for single
198
Hg
+
ions to be
about 0.15 s [44] and for single
9
Be
+
ions to be about
1 ms [45]. These relatively short times are, so far, unex-
plained; however, it might be possible to achieve much
longer times in the future (Sec. 4.1).
In the linear trap, the radial COM vibration frequency
v
r
must be made sufficiently higher than the axial COM
vibrational frequency
v
z
in order for the ions to be
collinear along the z axis of the trap. This configuration
will aid in addressing individual ions with laser beams
and will also suppress rf heating (Sec. 4.1.5). To prevent
zig-zag and other complicated shapes of the ion crystal,
we require
v
r
/
v
z
> 1 for two ions, and
v
r
/
v
z
> 1.55 for
three ions. For L > 3 ions, the critical ratio (
v
r
/
v
z
)
c
for
linear confinement has been estimated analytically [80,
81] yielding (
v
r
/
v
z
)
c
. 0.73L
0.86
[60]. Other estimates
are given in Refs. [82] [(
v
r
/
v
z
)
c
. 0.63L
0.865
] and [64]
[(
v
r
/
v
z
)
c
. 0.59L
0.885
]. An equivalent result is obtained
if we consider that as the potential is weakened in the
radial direction, ions in a long string which are spaced
by distance s
c
near the center of the string, will first
break into a zig zag configuration. At the point where
the ions break into a zig-zag, the net outward force from
neighboring ions is equal to the inward trapping force.
If we equate these forces, we obtain
v
2
r
=
7
8
0
z
(3)
q
2
ms
3
c
, (7)
where
z
is the Riemann zeta function. As an example,
for
9
Be
+
ions, m . 9 u (atomic mass units) and
s
c
=3mm, we must have
v
r
/2
p
> 7.8 MHz to keep the
ions along the axis of the trap.
The equilibrium spacing of a linear configuration of
trapped ions is not uniform; the middle ions are spaced
closer than the outlying ions, as is apparent in Fig. 1. The
separation of two ions is s
2
=2
1/3
s, where
s =(q
2
/4
p
«
0
m
v
2
z
)
1/3
is a length scale of ion-ion
spacings; the adjacent separation of three ions is
s
3
= (5/4)
1/3
s. For L large, estimates of the minimum
separation of the center ions are given by s
c
(L)
. 2sL
–0.56
[60], 2.018sL
– 0.559
[61], 2.29L
–0.596
[83], and
1.92sL
– 2/3
[ln(0.8L)]
1/3
[59, 81]. For typical trapping
parameters, the ion-ion separations are on the order of a
few mm and the spatial spread of the zero-point vibra-
tional wavepackets are on the order of 10 nm. Thus there
is negligible wavefunction overlap between ions and
quantum statistics (Bose or Fermi) play no role in the
spatial wavefunction of an array.
Of the 3L normal modes of oscillation in a linear trap,
we are primarily interested in the L modes associated
with axial motion because we will preferentially couple
to them with applied laser fields. A remarkable feature
of the linear ion trap is that the axial mode frequencies
are nearly independent of L[1,60,61,84]. For two ions,
the axial normal mode frequencies are at
v
z
and Ï3
v
z
;
for three ions they are
v
z
, Ï3
v
z
, and (5.8)
1/2
v
z
. For
L > 3 ions, the Lth axial normal mode can be deter-
mined numerically [60,61,84].
2.2 Ion Motional and Internal Quantum States
A single ion’s motion, or the COM mode of a collec-
tion, has a simple description when the ions are trapped
in a purely static potential, which is the case for the axial
motion in a Penning trap or the axial motion in the trap
of Fig. 1. We will assume that the trap potentials are
quadratic [Eqs. (1) and (2)]. This is a valid approxima-
tion when the amplitudes of motion are small, because
the local potential, expanded about the equilibrium point
of the trap, is quadratic to a good approximation. In this
case, motion is harmonic. An ion trapped in a pondero-
motive potential [Eq. (6)] can be described effectively as
a simple harmonic oscillator, even though the Hamilto-
nian is actually time-dependent, so no stationary states
exist. For practical purposes, the system can be treated
as if the Hamiltonian were that of an ordinary,
time independent harmonic oscillator [34,35,85–91]
although modifications must be made for laser cooling
[92]. The classical micromotion (the terms which vary
as cos
V
T
t and cos2
V
T
t in Eq. (4)] may be viewed, in the
quantum picture, as causing the ion’s wavefunction to
breathe at the drive frequency
V
T
. This breathing mo-
tion is separated spectrally from the secular motion [(at
frequencies
v
x
and
v
y
in Eq. (4)]. Since the operations
we will consider rely on a resonant interaction at the
secular frequencies, we will average over the compo-
nents of motion at the drive frequency
V
T
. Therefore, to
a good approximation, the pseudopotential secular
motion behaves as an oscillator in a static potential.
263
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
main consequence of the quantum treatment is that
transition rates between quantum levels (Eq. (18),
below) are altered [34, 35]; however, these changes can
be accounted for by experimental calibration. In any
case, for most of the applications discussed in this paper,
we will be considering the motion of the ions along the
axis of a linear Paul trap where this modification is
absent.
Therefore, the Hamiltonian describing motion of a
single ion (or a normal mode, such as the COM mode,
of a collection of ions) in the ith direction is given by
H
osc
=
"v
i
nˆ
i
, i[{x,y,z} , (8)
where nˆ
i
a
i
a
i
and a
i
and a
i
are the usual harmonic
oscillator raising and lowering operators and we have
suppressed the zero-point energy 1/2
"v
i
. The operator
for the COM motion in the z direction is given by
z = z
0
(a + a
) , (9)
where z
0
=(
"
/2m
v
z
)
1/2
is the spread of the zero-point
wavefunction and m is the ion mass. That is, z
0
=
(k0uz
2
u0l)
1/2
, where unl is the nth eigenstate (“number”
or Fock state) of the harmonic oscillator. For
9
Be
+
ions
in a trap where
v
z
/2
p
= 10 MHz, we have z
0
= 7.5 nm.
Therefore, a general pure state of motion for one mode
can be written, in the Schro¨dinger picture, as
C
motion
=
O
`
n=0
C
n
e
n
v
i
t
unl , (10)
where C
n
are complex and the unl are time-independent.
For applications to quantum logic, we will be interested
in motional states of the simple form
a
u0l +
b
exp(– i
v
i
t)u1l.
We will be interested in the situation where, at any
given time, we interact with only two internal levels of
an ion. This will be accomplished by insuring that the
internal states are nondegenerate and by using resonant
excitations to couple only two levels at a time. We will
find it convenient to represent a two-level system by its
analogy with a spin-1/2 magnetic moment in a static
magnetic field [93, 94]. In this equivalent representation,
we assume that a (fictitious) magnetic moment
m
=
m
M
S, where S is the spin operator (S = 1/2), is
placed in a (fictitious) magnetic field
B
= B
0
zˆ. The
Hamiltonian can therefore be written
H
internal
=
"v
0
S
z
, (11)
where S
z
is the operator for the z component of the spin
and
v
0
m
M
B
0
/
"
. Typically, the internal resonant
frequency will be much larger than any motional mode
frequency,
v
0
>>
v
z
.We label the internal eigenstates
uM
z
l = ul and ul representing “spin-up” and “spin-
down” respectively, and for convenience, will assume
m
M
< 0 so that the energy of the ul state is higher than
the ul state. A general pure state of the two-level system
is then given by
C
internal
= C
e
i
v
0
t
2
ul + C
e
i
v
0
t
2
ul , (12)
where uC
u
2
+ uC
u
2
= 1. Of course, the two-level system
really could be a S = 1/2 spin such as a trapped electron
or the ground state of an atomic ion with a single
unpaired outer electron and zero nuclear spin such as
24
Mg
+
.
2.2.1 Detection of Internal States
The applications considered below will benefit from
high detection efficiency of the ion’s internal states.
Unit detection efficiency has been achieved in experi-
ments on “quantum jumps” [95–99] where the internal
state of the ion is indicated by light scattering (or lack
thereof), correlated with the ion’s internal state. (More
recently, this type of detection has been used in spec-
troscopy so that the noise is limited by the fundamental
quantum fluctuations in detection of the internal state
[100]. In these experiments, detection is accomplished
with a laser beam appropriately polarized and tuned to
a transition that will scatter many photons if the atom is
in one internal state (a “cycling” transition), but will
scatter essentially no photons if the atom is in the other
internal state. If a modest number of these photons are
detected, the efficiency of our ability to discriminate
between these two states approaches 100 %. We note
that for a string of ions in a linear trap, the scattered light
from one ion will impinge on the other ions; this can
affect the detection efficiency since the scattered light
will, in general, have a different polarization.
The overall efficiency can be explained as follows.
Suppose the atom scatters N total photons if it is
measured to be in state ul and no photons if it is mea-
sured to be in state ul. In practice N will be limited by
optical pumping but can be 10
6
or higher [101]. Here,
we assume that N is large enough that we can neglect its
fluctuations from experiment to experiment. We
typically detect only a small fraction of these photons
due to small solid angle collection and small detector
efficiency. Therefore, on average, we detect n
d
=
h
d
N
photons where
h
d
<< 1 is the net photon detection
efficiency. If we can neglect background, then for each
experiment, if we detect at least one scattered photon,
we can assume the ion is in state ul. If we detect
no photons, the probability of a false reading, that is, the
probability the ion is in state ul but we simply did not
264
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
detect any photons, is given by P
N
(0) =
(1
h
d
)
N
. exp(–n
d
). For n
d
= 10, P
N
(0) . 4.5 3 10
–5
,
for n
d
= 100, P
N
(0) . 4 3 10
–44
. Therefore for n
d
> 10,
detection can be highly efficient.
Detection of ion motion can be accomplished directly
by observing the currents induced in the trap electrodes
[77–79, 102–104]. However, the sensitivity of this
method is limited by electronic detection noise. Because
the detection of internal states can be so efficient, mo-
tional states can be detected by mapping their properties
onto internal states which are then detected (Sec. 3.2).
2.3 Interaction With Additional Applied Electro-
magnetic Fields
2.3.1 Single Ion, Single Applied Field, Single Mode
of Motion
We first consider the situation where a single, period-
ically varying, (classical) electromagnetic field propa-
gating along the z direction is applied to a single trapped
ion which is constrained to move in the z direction in a
harmonic well with frequency
v
z
. We consider situa-
tions where fields resonantly drive transitions between
internal or motional states and when they drive transi-
tions between these states simultaneously (entangle-
ment). If we assume that the internal levels
are coupled by electric fields, then the interaction
Hamiltonian is
H
I
=–
m
d
? E(z, t) , (13)
where
m
d
is the electric dipole operator for the internal
transition and E is from a uniform wave propagating
along the z direction and polarized in the x direction,
E = E
1
xˆcos(kz –
v
t +
f
), where
v
is the frequency, k is
the wavevector 2
p
/
l
,and
l
is the wavelength. In the
equivalent spin-1/2 analog, we assume that a traveling
wave magnetic field propagates along the z direction, is
polarized in the x direction [B = B
1
xˆ cos(kz
v
t +
f
)],
and interacts with the fictitious spin (
m
=
m
M
S).
Therefore, for the spin analog, Eq. (13) is replaced by
H
I
=–
m
? B(z, t)
=
"V
(S
+
+ S
)(e
i(kz
v
t+
f
)
+e
i(kz
v
t+
f
)
) , (14)
where
"V
m
M
B
1
/4 (or
m
d
E
1
/4 for an electric
dipole), S
+
S
x
+ iS
y
, S
S
x
iS
y
and z is given by
Eq. (9). We will assume that the lifetimes of the levels
are long; in this case, the spectrum of the transitions
excited by the traveling wave is well resolved if
V
is
sufficiently small.
It will be useful to transform to an interaction picture
where we assume H
0
= H
internal
+ H
osc
and V
interaction
= H
I
.
In this interaction picture, if we make the rotating-wave
approximation (neglecting exp(6i(
v
+
v
0
)t) terms),
the wavefunction can be written
C
=
O
M
z
=,
O
`
n=0
C
M
z
,n
(t)uM
z
lunl , (15)
where uM
z
l and unl are the time-independent internal
and motional eigenstates. In general, this wavefunction
will be entangled between the two degrees of freedom;
that is, we will not be able to write the wavefunction as
a product of internal and motional wavefunctions.
We h ave H H'
I
= U
0
(t)H
1
U
0
(t) where U
0
(t)=
exp( i(H
0
/
"
)t), resulting in
H'
I
=
"V
S
+
exp(i[
h
(ae
i
v
z
t
+ a
e
i
v
z
t
)–
d
t +
f
])
+ h.c.(
d
v
v
0
) , (16)
where
h
kz
0
is the Lamb-Dicke parameter and
S
+
, S
, a,anda
are time independent. We will be
primarily interested in resonant transitions, that is,
where
d
=
v
z
(n' n) where n' and n are integers. How-
ever, since we want to consider nonideal realizations, we
will assume
d
=(n' n)
v
z
+
D
where u
D
u <<
v
z
,
V
.If
we can neglect couplings to other levels (see Sec. 4.4.6),
transitions are coherently driven between levels u, n l
and u, n' l and the coefficients in Eq. (15) are given by
Schro¨dinger’s equation i
"
dC
/
t = H'
I
C
to be
C
˙
,n'
=–i
(1 + un' nu)
e
i(Dt
f
)
V
n',n
C
,n
,
C
˙
,n
=–i
(1 un' nu)
e
i(Dt
f
)
V
n',n
C
,n'
, (17)
where
V
n',n
is given by [105, 106]
V
n',n
V
ukn'u e
i
h
(a+a
)
unlu
=
V
exp[–
h
2
/2](n
<
!/n
>
!)
1/2
h
un'n|
L
n
<
un'nu
(
h
2
) , (18)
where n
<
(n
>
) is the lesser (greater) of n' and n,andL
a
n
is the generalized Laguerre polynomial
L
a
n
(X)=
O
n
m=0
(–1)
m
S
n +
a
n m
D
X
m
m!
. (19)
Since we will be particularly interested in small values
of n and
a
, for convenience, we list a few values of
L
a
n
(X).
265
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
L
0
0
(X)=1,L
0
1
(X)=1–X, L
0
2
(X)1–2X +
X
2
2
, L
0
3
(X)=1–3X +
3
2
X
2
1
6
X
3
L
1
0
(X)=1,L
1
1
(X)=2–X, L
1
2
(X)=3–3X +
1
2
X
2
, L
1
3
(X)=4–6X +2X
2
1
6
X
3
L
0
2
(X)=1,L
2
1
(X)=3–X, L
2
2
(X)=6–4X +
1
2
X
2
, L
2
3
(X)=10–10X +
5
2
X
2
1
6
X
3
(20)
Equations (17) can be solved using Laplace transforms.
The solution shows sinusoidal “Rabi oscillations”
between the states u, n'l and u, nl, so over the
subspace of these two states we have
C
(t)=
e
i
D
2
t
F
cos
S
X
n',n
2
t
D
+ i
D
X
n',n
sin
S
X
n',n
2
t
DG
–2i
V
n',n
X
n',n
e
i
S
D
2
t
f
p
2
un'nu
D
sin
S
X
n',n
2
t
D
–2i
V
n',n
X
n',n
e
+i
S
D
2
t
f
p
2
un'nu
D
sin
S
X
n',n
2
t
D
e
i
D
2
t
F
i
D
X
n',n
sin
S
X
n',n
2
t
D
+ cos
S
X
n',n
2
t
DG
C
(0),
(21)
_ _
_
_
where X
n',n
(
D
2
+4
V
2
n',n
)
1/2
,
D
=
v
v
0
–(n' n)
v
z
, and
C
is given by
C
= C
,n
u, nl + C
,n'
u,n'l =
F
C
,n'
C
,n
G
. (22)
For the resonance condition
D
= 0, Eq. (21) simplifies
to
C
(t)=
3
cos
V
n',n
t
i e
i[
f
+
p
2
un'nu]
sin
V
n',n
t
ie
i[
f
+
p
2
un' nu]
sin
V
n',n
t
cos
V
n',n
t
4
C
(0) . (23)
then
h
<< 1, but the converse is not necessarily true. If
the Lamb-Dicke criterion is satisfied, we can evaluate
V
n',n
to lowest order in
h
to obtain
V
n',n
=
V
n,n'
=
Vh
un' nu
(n
>
!/n
<
!)
1/2
(un' nu!)
–1
. (24)
We will be primarily interested in three types of transi-
tions—the carrier (n' = n), the first red sideband
(n' = n –1), and the first blue sideband (n' = n +1)
whose Rabi frequencies, in the Lamb-Dicke limit, are
given from Eq. (24) by
V
,
h
n
1/2
V
, and
h
(n +1)
1/2
V
respectively.
In general, the Lamb-Dicke limit is not rigorously
satisfied and higher order terms must be accounted for
in the interaction [16,21,106]. As a simple example,
suppose
C
(0) = ulunl and we apply radiation at
the carrier frequency (
d
= 0). From Eq. (23), the wave-
function evolves as
When the atom starts in an eigenstate, for each value of
n' n, the phase factor
f
+
p
un'– nu/2 can be chosen
arbitrarily for the first application of H
I
; however once
chosen, it must be kept track of if subsequent applica-
tions of H
I
are performed on the same ion. For conve-
nience, we can choose it to be zero, although in most of
what follows we will include a phase factor as a re-
minder that we must keep track of it. In these expres-
sions, we assume
V
n',n
to be constant during a given
application time t; this condition can be relaxed as
discussed in Sec. 4.3.2. A special case of interest is
when the Lamb-Dicke criterion, or Lamb-Dicke limit, is
satisfied. Here, the amplitude of the ion’s motion in the
direction of the radiation is much less than
l
/2
p
which
corresponds to the condition k
C
motion
uk
2
z
2
u
C
motion
l
1/2
<< 1. This should not be confused with the less restric-
tive condition where the Lamb-Dicke parameter is less
than 1 (
h
<< 1); if the Lamb-Dicke criterion is satisfied,
266
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
C
(t) = cos
V
n,n
tu, nl ie
i
f
sin
V
n,n
tu, nl . (25)
For n =0,wehave
V
0,0
=
V
exp(–
h
2
/2). The exponential
factor in this expression is the Debye-Waller factor
familiar from studies of x-ray scattering in solids; for a
discussion in the context of trapped atoms see Ref. [106]
and Sec. 4.4.5. This factor indicates that the matrix
element for absorption of a photon is reduced due to the
averaging of the electromagnetic wave (averaging of the
e
ikz
factor in Eq. (14)) over the spread of the atom’s
zero-point wavefunction.
As a second simple example, we consider
C
(0) = u,nl and
d
=+
v
z
(first blue sideband). Equa-
tion (23) implies
C
(t) = cos
V
n+1,n
tu, nl +e
i
f
sin
V
n +1,n
tu, n +1l. (26)
At any time t Þ m
p
/(2
V
n+1,n
)(m an integer),
C
is an
entangled state between the spin and motion. If the
excitation is left on continuously, the atom sinusoidally
oscillates between the state u, nl and u, n +1l. This
oscillation has been observed in Ref. [21] and is repro-
duced in Fig. 2.
When the Lamb-Dicke confinement criterion is met
and when the radiation is tuned to the red sideband
(
d
=–
v
z
), we find (choosing
f
=–
p
/2)
H
I
=
"hV
(S
+
a + S
a
) . (27)
This Hamiltonian is the same as the “Jaynes-Cummings
Hamiltonian” [107] of cavity QED [43], which
describes the coupling of a two-level atom to a single
mode of the (quantized) radiation field. The problem we
have described here, the coupling of a single two-level
atom to the atom’s (harmonic) motion is entirely
analogous; the difference is that the harmonic oscillator
associated with a single mode of the radiation field in
cavity QED is replaced by that of the atom’s motion.
The suggestion to realize this type of Hamiltonian (in
the context of cavity-QED) with a trapped ion was
outlined in Refs. [2], [3], and [7]; however its use was
already employed in the g-2 single electron experiments
of Dehmelt [108].
Driving transitions between the ul and ul states will
create entangled states between the internal and
motional states since, in general, the Rabi frequency
will depend on the motional states (Eq. (18)). This
“conditional dynamics,” where the dynamics of one
system is conditioned on the state of another system,
provides the basis for quantum logic (Sec. 3.3).
In this section, we have assumed that the atom inter-
acts with an electromagnetic wave (Eq. (14)), which
will usually be a laser beam. However, the essential
physics which gives rise to entanglement is that the
atom’s internal levels are coupled to its motion through
an inhomogeneous applied field. In the spin-1/2 analog,
the magnetic moment
m
couples to a magnetic field
B = B(z,t)xˆ , yielding the Hamiltonian
H
I
=–
m
x
B(z, t)=
m
x
F
B(z =0,t)+
B
z
U
z=0
z +
1
2
2
B
z
2
U
z=0
z
2
+...
G
, (28)
where, as above,
m
x
~ S
+
+S
and z is the position
operator. The key term is the gradient
B/
z. From
the atom’s oscillatory motion in the z direction, it expe-
riences, in its rest frame, a modulation of B at frequency
v
z
. This oscillating component of B can then drive the
spin-flip transition. As a simple example, suppose B is
static (but inhomogeneous along the z direction so that
Fig. 2. Experimental plot of the probability P
(t) of finding a single
9
Be
+
ion in the ul state after first preparing
it in the ulu0l state and applying the first blue side band coupling (Eq. (16), for
d
=+
v
z
) for a time t. If there
were no decoherence in the system, P
(t) should be a perfect sinusoid as indicated in Eq. (26). Decoherence
causes the signal to decay as discussed in Sec. 3.2.1. The solid line is a fit to an exponentially decaying sinusoid
as indicated in Eq. (43). Each point represents an average of 4000 observations [21].
267
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
B/
z Þ 0) and
v
z
is equal to the resonance frequency
v
0
of the internal state transition. In its reference frame,
the atom experiences an oscillating field due to the
motion through the inhomogeneous field. Since
v
=
v
0
,
this field resonantly drives transitions between the inter-
nal states. Because this term is resonant, it is the domi-
nant term in Eq. (28), so H
I
.
m
x
(
B/
z)z ~ (S
+
+
S
)(a + a
) . S
+
a + S
a
, where the last equality ne-
glects nonresonant terms. If the extent of the atom’s
motion is small enough that we need only consider the
first two terms on the right hand side of Eq. (28), H
I
is
given by the Janes-Cummings Hamiltonian (Eq. 27)).
This Hamiltonian is also obtained if B is sinusoidally
time varying (frequency
v
), we satisfy the resonance
condition
d
=
v
v
0
=–
v
z
, and we make the rotating-
wave approximation. This situation was realized in the
classic electron g-2 experiments of Dehmelt, Van Dyck,
and coworkers to couple the spin and cyclotron motion
[108]. Higher-order sidebands are obtained by consider-
ing higher order terms in the expansion of Eq. (28).
One reason to use optical fields is that the field gradi-
ents (for example,
/(
z)[e
ikz
]=ike
ikz
) can be large
because of the smallness of
l
. Stated another way,
single-photon transitions between levels separated by rf
or microwave transitions, which are driven by plane
waves, may not be of interest because k is small (
l
large)
and exp(ikz) . 1 which implies
/(
z)[e
ikz
] . 0. This
makes interactions which couple the internal and exter-
nal states as in Eqs. (27) and (28) negligibly small. This
is not a fundamental restriction because electrode struc-
tures whose dimensions are small compared to the wave
length can be used to achieve much stronger gradients
than are achieved with plane waves. Microwave or rf
transitions can also be driven by using stimulated-
Raman transitions as discussed in Sec. 2.3.3 below. A
second reason to use laser fields is they can be focused
so that, to a good approximation, they interact only with
a selected ion in a collection.
The unitary transformations of Eqs. (21) and (23)
form the basic operations upon which most of the
manipulations discussed in this paper are based. In this
section, they were used to describe transitions between
two states labeled ulunl and ulun'l. In what follows, we
will include other internal states of the atom which will
take on different labels; however, the transitions
between selected individual levels can still be described
by Eqs. (21) and (23). Sequences of these basic opera-
tions can be used to construct more complicated opera-
tions such as logic gates (Sec. 3.3).
2.3.2 State Dynamics Including Multiple Modes of
Motion
In what follows, we will generalize the interaction
with electromagnetic fields to consider motion in all 3L
modes of motion for L trapped ions. Here, as was as-
sumed by Cirac and Zoller [1], we consider that, on any
given operation, the laser beam(s) interacts with only the
jth ion; however, that ion will, in general, have compo-
nents of motion from all modes. In this case Eq. (14) for
the jth ion becomes
H
Ij
=
"V
(S
+j
+ S
j
)[e
i(k ? x
j
v
t +
f
j
)
+ h.c.] , (29)
where we now assume k has some arbitrary direction.
We will write the position operator of the jth ion (which
represents the deviation from its equilibrium position)
as
x
j
= u
j
xˆ+u
L + j
yˆ+u
2L + j
zˆ, j[{1, 2, . . . L} . (30)
We can express the u
j
in terms of normal mode coordi-
nates q
k
(k[{1, 2, . . . 3L}) through the matrix D
p
k
,by
the following relations [109]
u
p
=
O
3L
k =1
D
p
k
q
k
, q
k
=
O
3L
p =1
D
p
k
u
p
, q
k
q
k0
(a
k
+ a
k
),
(31)
where q
k
is the operator for the kth normal mode and a
k
and a
k
are the lowering and raising operators for the kth
mode. We have assumed that all normal modes are
harmonic, which is a reasonably good assumption as
long as the amplitude of normal mode motion is small
compared to the ion spacing. (For two ions, the axial
stretch mode’s frequency is approximately equal to
v
z
(Ï3– 9(a
z
/a
s
)
2
) where a
z
is the (classical) amplitude
of one ion’s motion for this mode and a
s
is the ion
spacing). Following the procedure of the last section,
we take H
0
to be the Hamiltonian of the jth ion’s inter-
nal states and all of the motional (normal) modes
H
0j
=
"v
0
S
zj
+
O
3L
k=1
"v
k
nˆ
k
, (32)
where nˆ
k
a
k
a
k
. In the interaction picture (and making
the rotating wave approximation), we have H'
Ij
=
U
0j
H
Ij
U
0j
where U
0j
= exp(– i(H
0j
/
"
)t, yielding
H'
Ij
=
"V
S
+j
exp
F
i
O
3L
k =1
h
j
k
(a
k
e
i
v
k
t
+ a
k
e
i
v
k
t
)–i(dt
f
j
)
G
+ h.c. , (33)
where
h
j
k
(k ? xˆD
j
k
+ k ? yˆ D
L+j
k
+ k ? zˆD
2L+j
k
)q
k0
. (For
the linear trap case, motion will be separable in the
268
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
x, y,andz directions and
h
j
k
will consist of one term.)
In this interaction picture, the wavefunction is given by
C
j
=
O
M
z
= ,
O
`
{n
k
}=0
C
j
M
z
,{n
k
}
(t)uM
z
l
j
u{n
k
}l , (34)
where the coefficients are slowly varying and u{n
k
}l are
the normal mode eigenstates (we have used the short-
hand notation {n
k
}=n
1
, n
2
,...,n
3L
). In analogy with
the previous section, we will be primarily interested in
a particular resonance condition, that is, where
d
.
v
k
(n'
k
n
k
)and
V
is sufficiently small that coupling to
other internal levels and motional modes can be ne-
glected. In this case, Eqs. (21) and (23) apply to the
subspace of states ul
j
un
k
l and ul
j
un'
k
l if we make the
definitions
C
=
C
j
= C
j
,n
k
ul
j
un
k
l + C
j
,n'
k
ul
j
un
'
k
l
=
F
C
j
,n'
k
C
j
,n
k
G
,
D
d
–(n
'
k
n
k
)
v
k
, (35)
and
X
j
n
k
',n
k
(
D
2
+4(
V
j
n
k
',n
k
)
2
)
1/2
,
V
j
n
k
',n
k
V
uk{n
p Þ k
}, n
'
k
u P
3L
l=1
e
i
h
j
l
(a
l
+ a
l
)
u{n
pÞk
}, n
k
lu.
(36)
The last expression is the Rabi frequency for particular
values of the {n
pÞ k
}. More likely, the other mode states
(pÞ k) will correspond to a statistical distribution; this
is discussed in Sec. 4.4.5. For the application to
quantum logic (Sec. 3.3) the COM mode appears to be
a natural choice since
h
j
k
will be independent of j. The
dependence of
h
j
k
on j for the other modes is not a
fundamental problem, but requires accurate bookkeep-
ing when addressing different ions. The values of
h
j
k
can
be obtained from the normal mode coefficients as
described by James [61].
2.3.3 Stimulated-Raman Transition
As indicated in the discussion following Eq. (28), we
want strong field gradients to couple the internal states
to the motion. If the internal state transition frequency
v
0
is small, one way we can achieve strong field gradi-
ents is by using two-photon stimulated-Raman transi-
tions [45, 48] through a third, optical level as indicated
in Fig. 3. In this case, as we outline below, the effective
Hamiltonian corresponding to that in Eq. (14) is
replaced by
H
I
=
"V
(S
+
+ S
)[e
i[(k
1
k
2
) ? x –(
v
L1
v
L2
)t +
f
]
+ h.c.] , (37)
where k
1
, k
2
and
v
L1
,
v
L2
are the wavevectors and
frequencies of the two laser beams and the resonance
condition between internal states corresponds to
u
v
L1
v
L2
u =
v
0
. Even if
v
0
is small compared to optical
frequencies, uk
1
k
2
u can correspond to the wavevector
of an optical frequency by choosing different directions
for k
1
and k
2
; this choice can thereby provide the desired
strong field gradients.
In Fig. 3 we consider that a transition is driven
between states ul and ul through state u3l by stimu-
lated-Raman transitions using plane waves. Typically,
we consider coupling with electric dipole transitions in
which case
E
i
=
«
ˆ
i
E
i
cos(k
i
? x
v
Li
t +
f
i
), i[{1, 2} . (38)
For simplicity, we assume laser beam 1 has a cou-
pling only between intermediate state u3l and state ul.
Similarly, laser beam 2 has a coupling only between
state u3l and state ul. Not shown in Fig. 3 are the
energy levels corresponding to the 3L motional modes.
Laser detunings are indicated in the figure, so
v
L1
–(
v
L2
+
d
)=
v
0
, and we assume
D
R
>>
d
,{
v
k
}
where {
v
k
} are the 3L mode frequencies. We will
Fig. 3. Schematic diagram relevant to stimulated-Raman transitions
between internal states ul and ul. Two plane wave radiation fields
couple to a third state u3l. The radiation fields are typically at laser
frequencies; they are characterized by frequencies and wavevectors
v
Li
and k
i
, i[{1, 2}. The couplings are typically described by electric
dipole matrix elements. For simplicity, we assume field 1 only cou-
ples states ul and u3l; and field 2 only couples states ul and u3l.In
this diagram, we do not show the additional energy level structure of
the 3 L modes of motion.
269
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
assume the Raman beams are focussed so that they
interact only with the jth ion. In the Schro¨dinger picture,
the wavefunction is written
C
j
=
O
M
z
= , ,3
O
`
{n
k
}=0
C
j
M
z
,{n
k
}
3 exp[– i(
v
M
z
+ n
1
v
1
+ n
2
v
2
+... n
3L
v
3L
)t]
3 uM
z
l
j
u{n
3L
}l , (39)
Since
D
R
is large, state u3l can be adiabatically elimi-
nated in a theoretical treatment (see, for example,
Refs. [32], [48], [110], and Sec. 4.4.6.2). If we assume
the difference frequency is tuned to a particular reso-
nance
d
=
v
k
(n'
k
n
k
), we can neglect rapidly varying
terms and obtain
C
Ù
j
,n
1
,...n
'
k
,...n
3L
= i
ug
2
u
2
D
R
C
j
,n
1
,...n
'
k
,...n
3L
i
V
j
n
k
' , n
k
C
j
,n
1
,...n
k
,...n
3L
,
C
Ù
j
,n
1
,...n
k
,...n
3L
= i
ug
1
u
2
D
R
C
j
,n
1
,...n
k
,...n
3L
i(
V
j
n
k
', n
k
)
*
C
j
,n
1
,...n
'
k
,...n
3L
,
(40)
where
V
j
n'
k
,n
k
g
*
1
g
2
D
R
kn'
k
ue
i
h
j
k
(a
k
+ a
k
)
un
k
l , (41)
where g
1
E
1
e ku
«ˆ
? ru3l exp(– i
f
1
)/(2
"
),
g
2
E
2
eku
«ˆ
? ru3lexp(– i
f
2
)/(2
"
),
h
j
k
(Dk ? xˆD
j
k
+
Dk ? yˆD
L+j
k
+ Dk ? zˆD
2L + j
k
)q
k0
and Dk k
1
k
2
. The
terms ug
2
u/
D
R
and ug
2
u/
D
R
are the optical Stark shifts of
levels u1l and u2l respectively. They can be eliminated
from Eqs. (40) by including them in the definitions of
the energies for the ul and ul states or, equivalently,
tuning the Raman beam difference frequency
d
to
compensate for these shifts. If the Stark shifts are equal,
both the ul and ul states are shifted by the same
amount, and there is no additional phase shift to be
accounted (Sec. 4.4.3). Equations (40) for stimulated-
Raman transition amplitudes are the same as for the
two-level system (Sec. 2.3.1) if we make the identifica-
tions
f
1
f
2
f
and Dk k. Although the
experiments can benefit from use of stimulated-Raman
transitions, for simplicity, we will assume single photon
transitions below except where noted.
Another advantage of using stimulated-Raman
transitions on low frequency transitions, as opposed to
single-photon optical transitions, is the difference
frequency between Raman beams can be precisely
controlled using an acousto-optic modulator (AOM) to
generate the two beams from a single laser beam. If the
laser frequency fluctuations are much less than
D
R
,
phase errors on the overall Raman transitions can be
negligible [111]. Other advantages (and some disadvan-
tages) are noted below.
3. Quantum-State Manipulation
3.1 Laser Cooling to the Ground State of Motion
As a starting point for all of the quantum-state manip-
ulations described below, we will need to initialize the
ion(s) in known pure states. Using standard optical
pumping techniques [112], we can prepare the ions in
the ul internal state. Laser cooling in the resolved side-
band limit [106, 113 ] can generate the un =0l motional
state with reasonable efficiency [44, 45]. This type of
laser cooling is usually preceded by a stage of
“Doppler” laser cooling [106,114,115] which cools the
ion to an equivalent temperature of about 1 mK. For
Doppler cooling, we have knˆ l $ 1, so an additional stage
of cooling is required.
Resolved sideband laser cooling for a single, harmon-
ically-bound atom can be explained as follows: For sim-
plicity, we assume the atom is confined by a 1-D
harmonic well of vibration frequency
v
z
. We use an
optical transition whose radiative linewidth
g
rad
is rela-
tively narrow,
g
rad
<<
v
z
(Doppler laser cooling applies
when
g
rad
$
v
z
). If a laser beam (frequency
v
) is inci-
dent along the direction of the atomic motion, the bound
atom’s absorption spectrum is composed of a “carrier”
at frequency
v
0
and resolved frequency-modulation
sidebands that are spaced by
v
z
, that is, at frequencies
v
0
+(n' n)
v
z
(Sec. 2.3). These sidebands in the spec-
trum are generated from the Doppler effect (like vibra-
tional substructure in a molecular optical spectrum).
Laser cooling can occur if the laser is tuned to a lower
(red) sideband, for example, at
v
=
v
0
v
z
. In this case,
photons of energy
"
(
v
0
v
z
) are absorbed, and sponta-
neously emitted photons of average energy
"v
0
R
return the atom to its initial internal state, where R
(
"
k)
2
/2m =
"v
R
is the photon recoil energy of the atom.
Overall, for each scattering event, this reduces
the atom’s kinetic energy by
"v
z
if
v
z
>>
v
R
, a condition
which is satisfied for ions in strong traps. Since
v
R
/
v
z
=
h
2
where
h
is the Lamb-Dicke parameter, this
simple form of sideband cooling requires that the Lamb-
Dicke parameter be small. For example, in
9
Be
+
,ifthe
recoil corresponds to spontaneous emission from the
313 nm 2p
2
P
1/2
2s
2
S
1/2
transition (typically used for
laser cooling),
v
R
/2
p
. 230 kHz. This is to be
270
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
compared to trap oscillation frequencies in some laser-
cooling experiments of around 10 MHz [45]. Cooling
proceeds until the atom’s mean vibrational quantum
number in the harmonic well is given by
knˆl
min
. (
g
/2
v
z
)
2
<< 1 [106,115,116].
In experiments, we find it convenient to use two-
photon stimulated Raman transitions for sideband cool-
ing [45, 117], but the basic idea for, and limits to,
cooling are essentially the same as for single-photon
transitions. The steps required for sideband laser cooling
using stimulated-Raman transitions are illustrated in
Fig. 4. This figure is similar to Fig. 3, except we include
the quantum states of the harmonic oscillator for one
mode of motion. Part (a) of this figure shows how, when
the ion starts in the ul internal state, a stimulated-
Raman transition tuned to the first red sideband
ulunl ulun –1l reduces the motional energy by
"v
z
. In part (b), the atom is reset to the ul internal state
by a spontaneous-Raman transition from a third laser
beam tuned to the ul u3l transition. We assume that
there is a reasonable branching ratio from state u3l to
state ul, so that even if the atom decays back to level ul
after being excited to level u3l, after a few scattering
events, the atom decays to state ul.If
v
R
<<
v
z
, step (b)
accomplishes the transition ulun –1l ulun –1l with
high efficiency. Steps (a) and (b) are repeated until the
atom is optically pumped into the ulu 0l state. When this
condition is reached, neither step (a) or (b) is active and
the process stops. In this simple discussion, we have
assumed the transition ulunl ulun –1l is accom-
plished with 100% efficiency. However since, in
general, the atom doesn’t start in a given motional state
unl, and since the Rabi frequencies (Eq. (18)) depend on
n, this process is not 100 % efficient; nevertheless, the
atom will still be pumped to the ulu0l state. The only
danger is having the stimulated-Raman intensities
and pulse time t adjusted so that for a particular n,
V
n -1,n
t = m
p
(m an integer), in which case the atom is
“trapped” in the ulunl level. This is avoided by varying
the laser beam intensities from pulse to pulse; one par-
ticular strategy is described in Ref. [17].
So far, laser cooling to the un =0l state has been
achieved only with single ions [44, 45]; therefore an
immediate goal of future work is to laser cool a collec-
tion of ions (or, at least one mode of the collection) to the
zero-point state. Cooling of any of the 3L modes of
motion of a collection of ions should, in principle, work
the same as cooling of a single ion. To cool a particular
mode, we tune the cooling radiation to its first lower
sideband. If we want to cool all modes, sideband cooling
Fig. 4. Schematic diagram relevant to laser cooling using stimulated-Raman transitions. In (a), we
show that when
v
L1
v
L2
=
v
0
v
z
, stimulated-Raman transitions can accomplish the transition ul
un l ulun –1l. In the figure, the transition for n = 2 is shown. In (b), spontaneous-Raman transtions,
accomplished with radiation tuned to the ul u3l transition, pumps the atom back to the ul state,
thereby realizing the transition ulun –1l ulun –1l. When atomic recoil can be neglected, one
application of steps (a) and (b) reduces the atom’s motional energy by
"v
z
unless n = 0, in which case
the atom is in it’s motional ground state.
271
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
must be cycled through all 3L modes more than once, or
applied to all 3L modes at once, since recoil will heat all
modes. For the COM mode, the cooling is essentially
the same as cooling a single particle of mass Lm;how-
ever, the recoil energy upon re-emission is distributed
over the 3L–1 other modes. Other methods to prepare
atoms in the un =0l state are discussed in Refs. [5], [10],
[12]. Morigi et al. [118] show that it is not necessary to
satisfy the condition
v
R
<<
v
z
(
h
<< 1) to achieve
cooling to n =0.
3.2 Generation of Nonclassical States of Motion of
A Single Ion
We begin with a discussion of the generation of non-
classical motional states of a single trapped ion. This
seems appropriate because the other applications
discussed in this paper incorporate similar techniques.
Much of the original interest in nonclassical states
of mechanical motion grew out of the desire to make
sensitive detectors of gravitational waves using (macro-
scopic) mechanical resonators [119, 120]. For example,
parametric amplification of mechanical harmonic oscil-
lations can lead to quantum mechanical squeezing of the
oscillation. In the meantime, nonclassical states of the
radiation field were observed [43]. The close relation-
ship of these two problems was pointed out above: in
quantum optics, H
osc
of Eq. (8) represents a single mode
of the radiation field, and H
I
of Eq. (14) represents the
coupling between the (quantized) field and atom. The
nonclassical states of motion considered here, such as
squeezed states, are the direct analogs of the nonclassi-
cal photon states in quantum optics. They appear to be
of intrinsic interest because, as in cavity QED, they
allow the rather complicated dynamics of the simple
quantum system [described by the Hamiltonian in
Eq. (27)] to be studied. Before discussing some methods
to create nonclassical states, we consider one method
for analyzing them.
3.2.1 Population Analysis of Motional States
As described below in this section, from the ulu0l
state, it is possible to coherently create states of the form
ul
C
motion
where
C
motion
is given by Eq. (10). One way
we can analyze the motional state created is as follows
[21]: To the state ul
C
motion
, we apply radiation on the
first blue sideband (n'=n + 1 in Eq. (23)) for a time
t
.
We then measure the probability P
(
t
) that the ion is in
the ul internal state. In the experiments of Meekhof et
al. [21], the internal state ul is the 2s
2
S
1/2
(F =2,
M
F
= 2) state of
9
Be
+
,andul corresponds to the 2s
2
S
1/2
(1,1) state as shown in Fig. 5. The ul state is detected
by applying nearly resonant
s
+
-polarized laser radiation
between the ul and
2
P
3/2
(F =3,M
F
= 3) energy levels.
Because the only decay channel of the
2
P
3/2
(F =3,
M
F
= 3) state is back to the ul state, this is a cycling
transition, and detection efficiency is near 1 (Sec. 2.2.1).
The experiment is repeated many times for each value of
t
, and for a range of
t
values. We find
P
(
t
)=
1
2
S
1+
O
`
n =0
P
n
e
g
n
t
cos(2
V
n +1,n
t
)
D
, (42)
where P
n
uC
,n
u
2
is the probability of finding the ion in
state ulunl. The phenomenological decay constants
g
n
are introduced to model decoherence that occurs during
the application of the blue sideband. The measured sig-
nal P
(
t
) can be inverted (Fourier cosine transform),
allowing the extraction of the probability distribution of
vibrational state occupation P
n
.
3.2.2 Fock States
In Fig. 2, we show an experimental plot [21] of the
probability P
(
t
) of finding the ion in the ul internal
state after first preparing it in
the ulu0l state, and applying the first blue sideband
for a time
t
. From Eq. (23), we would expect
P
(
t
) = cos
2
V
1,0
t
; however, we clearly see the effects
Fig. 5. Hyperfine levels of the 2s
2
S
1/2
ground state of
9
Be
+
in a weak
magnetic field (not to scale). The energy levels are designated by
horizontal lines. Above the lines, the levels are represented by atomic
physics labels (F, M
F
) where F is the total angular momentum (elec-
tron plus nuclear angular momentum) and M
F
is the projection of the
angular momentum along the magnetic field axis. The separation of
Zeeman substates in the different F manifolds is approximately equal
to 0.7 3 10
10
B
0
Hz where B
0
is expressed is teslas. The separation of
the F = 1 and F = 2 manifolds is approximately 1.25 GHz at B
0
=0.
For simplicity of notation, in most of the paper we make the identifi-
cations uF =2,M
F
=2l ul, u1, 1l ul, u2, 0l uauxl.
272
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
of some decoherence process which we can represent
adequately by the first term in Eq. (42)
P
(
t
)=
1
2
S
1+e
g
0
t
cos2
V
1,0
t
D
. (43)
In this experiment, we think the decoherence is not
simply caused by fundamental (radiative) decoherence
but has contributions from fluctuations in laser power
(which cause fluctuations in
V
1,0
), fluctuations in trap
drive voltage V
0
(which cause fluctuations in
v
z
), and
fluctuations in the ambient magnetic field (which cause
fluctuations in
v
0
.
Neglecting for the moment the effects of decoher-
ence, we see that for times
t
= m
p
/(2
V
1,0
)(m an in-
teger), the ion is in a nonentangled state (ulu0l or
ulu1l). Therefore, if the ion starts in the ulu 0l state, we
can prepare the atom in the ulu1l state (the un =1l Fock
state) by applying the blue sideband for a time
t
=
p
/2
V
1,0
, a so called Rabi
p
pulse. For other times,
the ion is in an entangled state given by Eq. (26). This
operation and the analogous operation on the first red
sideband will form key elements of quantum logic using
trapped ions.
We can generate higher-n Fock states of motion by a
sequence of similar operations. For example, to generate
the ulu2l state, we start in the ulu0l state, apply a
p
pulse on the first blue sideband, followed by a
p
pulse
on the first red sideband. This leads to the sequence
ulu0l ulu1l ulu2l (neglecting overall phase fac-
tors). In a similar fashion, Fock states up to un =16l have
been created [21]. Other methods for creating Fock
states have been suggested in Refs. [5], [8], [12], and
[13].
3.2.3 Coherent States
We can also create coherent states of motion; these
states are closest in character to classical states of mo-
tion. This can be accomplished if the atom is subjected
to a spatially uniform classical force, or any force
derived from a potential f(t) ? z, where f is a real
c-number vector. For an ion which starts in the un =0l
state, this force creates a displacement leading to a co-
herent state u
a
l defined by a u
a
l =
a
u
a
l where
a
is a
complex number [121]. This classical force can be real-
ized by applying an electric field which oscillates at
frequency
v
z
. For example, if we apply a (classical)
electric field E(t)=z
ˆ
E
z
sin(
v
t–
f
), the corresponding
interaction Hamiltonian (in the interaction frame for the
motion) is given by
H
I
=–qE
z
z
0
(ae
i
v
z
t
+ a
e
i
v
z
t
) sin(
v
t
f
) . (44)
If we express the motional wavefunction as in Eq. (10),
Schro¨dinger’s equation yields for the coefficients of the
wavefunction
C
Ù
n
=
V
1
ÏnC
n–1
V
*
1
Ïn +1C
n +1
, (45)
where
V
1
qE
z
z
0
e
i
f
/(2
"
). Equivalently, in the inter-
action picture for the motion, the Hamiltonian of Eq.
(44) leads to the evolution operator
U(t)=e
[(
V
1
t)a
–(
V
1
t)*a]
= D(
V
1
t) , (46)
where D is the displacement operator with argument
V
1
t
[121].
We can achieve the same evolution if we superim-
pose two traveling wave fields which drive stimulated-
Raman transitions between different unl levels of the
same internal state, and we make the difference fre-
quency between the Raman beams equal to the trap
oscillation frequency. For example, assume the ion is
subjected to two lasers fields given by Eq. (38), where
v
L1
v
L2
.
v
z
<<
v
0
. The dynamics can be obtained
following the analysis in Sec. 2.3.3, except we replace
level ulunl (ulun'l) with level uglunl(uglun'l where ugl
can be any ground state which has a matrix element
with level u3l . For the coefficients of Eq. (10), we find
C
Ù
n
= i
F
ug
1
u
2
+ ug
2
u
2
D
R
C
n
O
`
n'=0
knu
V
exp[i
h
(ae
i
v
t
+ a
e
i
v
t
)–i(
v
L1
v
L2
)t]+h.c.un'lC
n'
G
, (47)
where
V
g
*
1
g
2
/
D
R
, g
i
qE
1
kgu
«ˆ
i
? r u3l exp(– i
f
i
)/
(2
"
),
h
(k
1
k
2
) ? zˆz
0
, and
D
R
v
3g
v
L1
. The first
term on the right side of Eq. (47) corresponds to a Stark
shift of level g; this Stark shift can be absorbed into the
definition of the ground state energy (see, for example,
Sec. 4.4.6.2). If this is done, the same equations for the
C
n
[Eqs. (47)] are obtained from the Hamiltonian (in the
oscillator interaction picture)
H'
I
=
"V
exp
S
i[
h
(ae
–i
v
z
t
+ a
e
i
v
z
t
)–(
v
L1
v
L2
)t]
D
+ h.c. , (48)
In the Lamb-Dicke limit, if we choose the resonance
condition where
v
L1
v
L2
=
v
z
and assume
V
<<
v
z
,
Eqs. (47) are the same as Eqs. (45) where
V
1
=
hV
.
Therefore, in the Lamb-Dicke limit, the applied laser
fields act like a uniform oscillating electric field which
oscillates at frequency
v
L1
v
L2
. This can be under-
stood if we consider that the two laser fields give rise
to an optical dipole force which is modulated in such
a way to resonantly excite the ion motion. To see
273
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
this, assume for simplicity that, in Eq. (38), E
1
= E
2
= E
0
,
«ˆ
1
=
«ˆ
2
=
«ˆ
,andk
1
k
2
= z
ˆ
uDku. It is useful to write the
total electric field as
E = E
1
+ E
2
=
«
ˆ2E(t)
3 cos((k
1
+ k
2
) ? z/2
v
t +
f
) , (49)
where
v¯
(
v
L1
+
v
L2
)/2, f (
f
1
+
f
2
)/2, and E(t)isa
slowly varying function
E(t)=E
0
cos
S
Dk
2
z
v
L1
v
L2
2
t +
f
1
f
2
2
D
. (50)
On a time scale long compared to 1/
D
R
but short com-
pared to 1/(
v
L1
v
L2
), the atom experiences a nonreso-
nant electric field of amplitude E(t) which is nearly
constant in time. If we consider coupling of this electric
field between the ground state and state u3l (for example,
see Sec. 4.4.6.2), this electric field leads to a spatially-
dependent Stark shift of the ground state equal to
D
E
Stark
=–4
"
ug(z,t)u
2
/
D
R
where
ug(z,t)u
2
= ugu
2
cos
2
S
Dk
2
z
v
L1
v
L2
2
t +
f
1
f
2
2
D
,
g
qE
0
kgu
«ˆ
? ru3le
i
f
2
"
. (51)
This Stark shift leads to an optical dipole force
122–124] F
z
=–
(
D
E
Stark
)/
z. On a longer time scale,
this dipole force is modulated at frequency
v
L1
v
L2
which can resonantly excite the ion’s motion when
(
v
L1
v
L2
)=
v
z
. This leads to Eqs. (45). When k
1
and
k
2
are both directed along the z axis (but in opposite
directions), the dipole force potential can be viewed as
a “moving standing wave” in the z direction which slips
over the ion and whose accompanying dipole force reso-
nantly excites the ion’s motion [125]. Both methods
have been used to excite coherent states in Refs. [21] and
[47]. Other methods for generating coherent states are
suggested in Refs. [4] and [48].
In the experiment of Ref. [47], a dipole force oscillat-
ing at the ion oscillation frequency was created with
particular polarizations of the laser fields. This led to a
force which was dependent on the ion’s internal state,
enabling the generation of entangled “Schro¨dinger-cat”
states of the form
C
=(ulu
a
e
i
f
l + ulu
a
e
i
f
l)/Ï2.
3.2.4 Other Nonclassical States
When (
v
L1
v
L2
)=2
v
, a similar analysis shows that
the moving standing wave potential discussed in the last
section has a component which acts like a parametric
excitation of the ion’s harmonic well at frequency 2
v
[21]. This can produce quantum mechanical squeezing
of the ion’s motion. Squeezing could also be achieved by
amplitude modulating U
0
at frequency 2
v
z
by a nonadi-
abatic change in the trap spring constant [48], or
through a combination of standing and traveling wave
laser fields [4]. A quantum mechanical treatment of the
motion in an rf trap shows the effects of squeezing from
the applied rf trapping fields [34, 35, 91]. More general
nonlinear effects in the interaction can lead to higher
“nonlinear coherent states” as discussed by de Matos
Filho and Vogel [126].
Other methods for generation of Schro¨dinger-cat like
states in ions are suggested in Refs. [8], [19], [29], [31],
[33], [36], and [38]. Additional nonclassical states are
investigated theoretically by Gou and Knight [23], Gou
et al. [37], and Gerry et al. [37]. Schemes which can
generate arbitrary states of the single-mode photon field
[127,128] can also be applied directly to generate
arbitrary motional states of a trapped atom and perform
quantum measurements of an arbitrary motional observ-
able [129]. A scheme which can generate arbitrary
entanglement between the internal and motional levels
of a trapped ion is discussed by Kneer and Law [130].
The procedure for analyzing motional states outlined
in Sec. 3.2.1 yields only the populations of the various
motional states and not the coherences. Coherences
must be verified separately [21, 47]. The most complete
characterization is achieved with a complete state
reconstruction or tomographic technique; a description
of how this has been implemented to measure the
density matrix or Wigner function for trapped atoms is
given in Leibfried et al. [131, 132]. These experiments
represent the first measurement of negative values of the
Wigner function in position-momentum space. Wigner
functions for free atoms have also been recently deter-
mined experimentally by Kurtsiefer et al. [133]. Other
methods for trapped atoms have been suggested in Refs.
[15], [22], [24], [27], [34], and [35]. These techniques
can also be extended to characterize entangled motional
states [39] and states which are entangled between the
motional and internal states.
3.3 Quantum Logic
Significant attention has been given recently to the
possibility of quantum computation. Although this field
is about 15 years old [134–138], interest has intensified
because of the discovery of algorithms, notably for
prime factorization [139–142], which could provide
dramatic speedup over conventional computers.
Quantum computation may also find other applications
274
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[142–152]. Schemes for implementing quantum compu-
tation have been proposed by Teich [153], Lloyd [143,
144], Berman et al. [154], DiVincenzo [155, 156], Cirac
and Zoller [1], Barenco et al. [157], Sleator and
Weinfurter [158], Pellizzari et al. [159], Domokos et al.
[160, 161] Turchette et al. [162], Lange et al. [163],
To¨rma¨ and Stenholm [164], Gershenfeld and Chuang
[165], Cory et al. [166], Privman et al. [167], Loss and
DiVincenzo [168], and Bocko et al. [169]. In this paper,
we focus on a scheme suggested by Cirac and Zoller [1],
which uses trapped ions. Since, in general, any quantum
computation can be composed of a series of single-bit
rotations and two-bit controlled-not operations
[140,155,156,170,171], we will focus our attention on
these operations.
In the parlance of quantum computation, we say that
two internal states of an ion can form a quantum bit or
“qubit” whose levels are labeled u0l and u1l or, equiva-
lently, ul and ul. Single-bit rotations on ion j can be
characterized by the transformation (Eq. (23) for n' = n)
R(
u
,
f
)(C
j
ul
j
+ C
j
ul
j
)
=
F
cos(
u
/2)
ie
i
f
sin(
u
/2)
ie
+i
f
sin(
u
/2)
cos(
u
/2)
G
3
F
C
j
C
j
G
. (52)
In the spin-1/2 model, this transformation is realized by
application of a magnetic field B
1
/2 which rotates at
frequency
v
0
and in the same sense as
m
which is
applied slong the direction xˆ cos
f
yˆ sin
f
in the rotat-
ing frame. This is equivalent to application of the field
B
1
xˆcos(kz –
v
t +
f
) in Eq. (14). In this expression,
u
=2
V
n,n
t is the angle of rotation about the axis of
this field. For
u
=
p
and
f
=0,R(
u
,
f
) is a logical
“not” operation (within an overall phase factor).
R(
p
/2, –
p
/2) (plus a rotation about z) is essentially a
Hadamard transform.
A fundamental two-bit gate is a controlled-not (CN)
gate [142,155,156,157]. This provides the transforma-
tion
u
«
1
lu
«
2
l u
«
1
lu
«
1
%
«
2
l , (53)
where
«
1
,
«
2
[{0, 1} and ! is addition modulo 2.
Although Eq. (53) is written in terms of eigenstates, the
transformation is assumed to apply to arbitrary superpo-
sitions of states u
«
1
lu
«
2
l. In this expression,
«
1
is the
called the control bit and
«
2
is the target bit. If
«
1
=0,the
target bit remains unchanged; if
«
1
= 1, the target bit
flips.
A spectroscopy experiment on any four-level quan-
tum system, where the level spacings are unequal, shows
this type of logic structure if we make the appropriate
labeling of the levels. For example, we could label these
four levels as in Eq. (53). If we tune radiation to the
u1, 0l u1, 1l resonance frequency and adjust its dura-
tion to make a
p
pulse, we realize the logic of Eq. (53).
Similarly, an eight-level quantum system with unequal
level spacings realizes a Toffoli gate [142], where the
flip of a third bit is conditioned upon the first two bits
being 1’s—and so on (see Sec. 5.2). This basic idea can
be applied to molecules composed of many interacting
spins such as in the proposals of Gershenfeld and
Chuang [165] and Cory et al. [166]. For quantum
computation to be most useful, however, we need to
perform a series of logic operations between an arbi-
trary number of qubits in a system which can be scaled
to large numbers, such as the scheme of Cirac and Zoller
[1].
Another type of fundamental two-bit gate is a phase
gate, which could take the form
u
«
1
lu
«
2
l e
i
1
«
2
u
«
1
lu
«
2
l . (54)
This type of conditional dynamics has been demon-
strated in the context of cavity QED [162,172] and for
a trapped ion [17] (step (1b) below).
The Cirac/Zoller scheme assumes that an array of
ions are confined in a common ion trap. The ions
are held apart from one another by mutual Coulomb
repulsion as shown, for example, in Fig. 1. They can be
individually addressed by focusing laser beams on the
selected ion. Ion motion can be described in terms of
normal modes of oscillation which are shared by all of
the ions; a particularly useful mode might be the COM
axial mode. When quantized, this mode can form the
“bus qubit” through which all gate operations are per-
formed. We first describe how logic is accomplished
between this COM mode qubit and the internal-state
qubit of a single trapped ion. In particular, the transfor-
mation in Eq. (53) has been realized for a single trapped
ion [17]. In that experiment, performed on a trapped
9
Be
+
ion, the control bit was the quantized state of one
mode of the ion’s motion (labeled the x mode). If
the motional state was un =0l, this was taken to be a
u
«
1
=0l state; if the motional state was un =1l, this was
taken to be a u
«
1
=1l state. The target states were two
ground-hyperfine states of the ion, the uF =2,M
F
=2l
and uF =1,M
F
=1l states, labeled ul and ul
(Fig. 5), with the identification here ul u
«
2
=0l
and ul u
«
2
=1l. Transitions between levels were
275
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
produced using two laser beams to realize stimulated-
Raman transitions. The wavevector difference k
1
k
2
was chosen to be aligned along the x direction. The CN
operation between these states was realized by applying
three pulses in succession:
(1a) A
p
/2 pulse (
V
t =
p
/4 in Eq. (25), where we
assume
V
0,0
=
V
1,1
=
V
) is applied on the carrier
transition. For a certain choice of initial phase,
this corresponds to the operator V
1/2
(
p
/2) of
Cirac and Zoller [1].
(1b) A 2
p
pulse is applied on the first blue sideband
transition between levels ul and an auxiliary
level uauxl in the ion (the uF =2,M
F
=0l level in
9
Be
+
; see Fig. 5). This operator is analogous to the
operator U
2,1
n
of Cirac and Zoller [1]. This opera-
tion provides the “conditional dynamics” for the
CN operation. It changes the sign of the ul
un =1l component of the wavefunction but leaves
the sign of the ulun =0l component of the wave-
function unchanged; that is, the sign change is
conditioned on whether or not the ion is in the
un =0l or un =1l motional state. Therefore, this
step is the phase gate of Eq. (54) with
f
=
p
,
where we make the identifications (
«
1
= 0,1)
(n = 0,1) and (
«
2
= 0,1) (internal state = ,).
(1c) A
p
/2 pulse is applied to the spin carrier transi-
tion with a 1808 phase shift relative to step (a).
This corresponds to the operator V
1/2
(–
p
/2) of
Cirac and Zoller [1].
Steps (1a) and (1c) can be regarded as two resonant
pulses (of opposite phase) in the Ramsey separated-field
method of spectroscopy [173]. If step (1b) is active
(thereby changing the sign of the ulun =1l component
of the wavefunction), then a state change (spin flip) is
induced by the Ramsey fields. If step (1b) is inactive,
step (1c) reverses the effect of step (1a).
Instead of the three pulses (1a – 1c above), a simpler
CN gate scheme between an ion’s internal and motional
states can be achieved with a single laser pulse, while
eliminating the requirement of the auxiliary internal
electronic level [174], as described below. These simpli-
fications can be important for several reasons:
(1) Several popular ion candidates, including
24
Mg
+
,
40
Ca
+
,
88
Sr
+
,
138
Ba
+
,
174
Yb
+
,
172
Yb
+
,and
198
Hg
+
,do
not have a third electronic ground state available
for the auxiliary level. These ions have zero nu-
clear spin with only two Zeeman ground states
(M
z
= ,). Although excited optical metastable
states may be suitable for auxiliary levels in some
of these ion species, use of such states places
stringent requirements on the frequency stability
of the exciting optical field to preserve coherence
(see Sec. 4.4.3).
(2) The elimination of an auxiliary ground state level
removes “spectator” internal atomic levels, which
can act as potential “leaks” from the two levels
spanned by the quantum bits (assuming negligible
population in excited electronic metastable
states). This feature may be important to the
success of quantum error-correction schemes
[142, 175–187] which can be degraded when
leaks to spectator states are present [188].
(Specific error-correction schemes for ions are
suggested in Refs. [182] and [187].)
(3) The elimination of the need for an auxiliary level
can dramatically reduce the sensitivity of a CN
quantum logic gate to external magnetic fields
fluctuations. It is generally impossible to find
three atomic ground states whose splittings are all
magnetic field insensitive to first order. However,
for ions possessing hyperfine structure, the tran-
sition frequency between two levels can be made
magnetic field independent to first order at partic-
ular values of an applied magnetic field (see Sec.
4.2.2).
(4) Finally, a reduction of laser pulses simplifies the
tuning procedure and may increase the speed of
the gate. For example, the gate realized in Ref.
[17] required the accurate setting of the phase and
frequency of three laser pulses, and the duration
of the gate was dominated by the transit time
through the auxiliary level.
The CN quantum logic gate can be realized with a single
pulse tuned to the carrier transition which couples the
states unlul and unlul with Rabi frequency
V
n,n
[see
Eqs. (18), (36), and (41)]. Considering only a single
mode of motion,
V
n,n
=
V
uknue
i
h
(a + a
)
unlu =
V
e
h
2
/2
+
n
(
h
2
) , (55)
where
+
n
(
h
2
) L
0
n
(
h
2
) (Eq. (18)). Specializing to the
un =0l and un =1l vibrational levels relevant to quantum
logic, we have
V
0,0
=
V
e
h
2
/2
,
V
1,1
=
V
e
h
2
/2
(1 –
h
2
) . (56)
The CN gate can be achieved in a single pulse by
setting
h
so that V
1,1
/V
0,0
=(2k + 1)/2m, with k and m
positive integers satisfying m > k $ 0. Setting
V
1,1
/
V
0,0
=2m/(2k + 1) will also work, with the roles of the u0l
and u1l motional states switched in Eq. (53). By driving
276
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
the carrier transition for a duration
t
such that
V
1,1
t
=(k + 1/2)
p
,ora“
p
-pulse” (mod 2
p
)onthe
unl = u1l component, this forces
V
0,0
t
= m
p
. Thus the
states ulu1l and ulu1l are swapped, while the states
ulu0l and ulu0l remain unaffected. The net unitary
transformation, in the {0, 0, 1, 1} basis is
100 0
010 0
.
000 ie
i
f
(– 1)
km
00ie
i
f
(–1)
km
0 (57)
This transformation is equivalent to the reduced CN of
Eq. (53), apart from phase factors which can be elimi-
nated by the appropriate settings of the phase of subse-
quent logic operations [157].
The “magic” values of the Lamb-Dicke parameter
which allow the above transformation satisfy
+
1
(
h
2
)=1–
h
2
=(2k + 1)/2m, and are tabulated in
Table I of Ref. [174] for the first few values. [For rf
(Paul) trap confinement along the COM motional mode,
the Rabi frequencies of Eqs. (55) and (56) must be
altered to include effects from the micromotion at the rf
drive frequency
V
T
. In the pseudopotential approxima-
tion (
v
<<
V
T
), this correction amounts to replacing the
Lamb-Dicke parameter
h
in this paper by
h
[1 –
v
/
(2Ï2
V
T
)], as pointed out by Bardroff et al. [34, 35].
However, there is no correction if the COM motional
mode is confined by static fields (such as the axial COM
mode of a linear trap.)] It may be desirable for the
reduced CN gate to employ the unl = u2l or unl = u3l state
instead of the unl = u1l state for error-correction of
motional state decoherence [182]. In these cases, the
“magic” Lamb-Dicke parameters satisfy
+
2
(
h
2
)=
1–2
h
2
+
h
4
/2=(2k + 1)/2m for quantum logic with
unl = u0l and u2l,or
+
3
(
h
2
)=1–3
h
2
+3
h
4
/2 –
h
6
/6 =
(2k + 1)/2m for quantum logic with unl = u0l and u3l.
This scheme places a more stringent requirement on
the accuracy of
V
and
h
, roughly by a factor of m.In
the two-photon Raman configuration (Sec. 2.3.3), the
Lamb-Dicke parameter
h
=|
D
kuz
0
can be controlled by
both the frequency of the trap (appearing in z
0
) and by
the geometrical wavevector difference
D
k of the
two Raman beams. Accurate setting of the Lamb-Dicke
parameter should therefore not be difficult. Both
CN-gate schemes are sensitive to excitation in other
modes as discussed in Sec. 4.4.5.
The CN operations between a motional and internal
state qubit described above can be incorporated to
provide an overall CN operation between two ions in a
collection of L ions. Here, we choose the particular ion
oscillator mode to be a COM mode of the collection.
Specifically, to realize a controlled-not C
ˆ
c,t
between two
ions c = control bit, t = target bit), we first assume the
COM mode is prepared in the zero-point state. The
initial state of the system is therefore given by
C
=
S O
M
1
= ,
O
M
2
= ,
...
O
M
L
= ,
C
M
1
, M
2
,...M
L
3 uM
1
l
1
uM
2
l
2
...uM
L
l
L
D
u0l . (58)
C
ˆ
c,t
can be accomplished with the following steps:
(2a) Apply a
p
-pulse on the red sideband of ion c.
This accomplishes the mapping (
a
ul
c
+
b
ul
c
)u0l ul
c
(
a
u0l +
b
u1l, and corresponds to
the operator U
1,0
m
of Cirac and Zoller [1]. We note
that in the NIST experiments [17], we prepare the
state (
a
ul + bul)u0l from the ul|0l state using
the carrier transition. We can then implement the
mapping (
a
ul + bul)u0l ul(
a
u0l +
b
u1l)by
applying a
p
-pulse on the red sideband. This is
the “keyboard” operation for preparation of arbi-
trary motional input states for the CN gate of
steps 1a – 1c above. Analogous mapping of inter-
nal state superpositions to motional state super-
positions are reported in Refs. [47], [131], and
[132].
(2b) Apply the CN operation (steps 1a 1c or, the
single carrier pulse described above) between the
COM motion and ion t.
(2c) Apply the inverse of step (2a).
Overall, C
ˆ
c,t
provides the unitary transformation (in the
{ul
c
ul
t
, ul
c
ul
t
, ul
c
ul
t
, ul
c
ul
t
} basis)
C
ˆ
c,t
=
1000
0100
0001
0010
(59)
which is the desired logic of Eq. (53). Effectively, C
ˆ
c,t
works by mapping the internal state of ion c onto the
COM motion, performing a CN between the motion and
ion t, and then mapping the COM state back onto ion c.
The resulting CN between ions c and t is the same as the
CN described by Cirac and Zoller [1], because the oper-
ations V
1/2
(
u
) and U
1,0
m
commute.
__
_ _
__
__
277
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
A third possibility, which also uses only one internal
state transition on each ion, is the following. We employ
two nondegenerate motional modes, which we label
here as 1 and 2. These might be the COM modes in two
different directions. We first map the internal state infor-
mation from two qubits j and k onto the separate
motional modes (which are both initially in the un =0l
1
un =0l
2
zero-point state). This can be accomplished as
described in step (2a) above. We then apply a condi-
tional phase gate (Eq. (54) with
f
=
p
) to the two
motional modes. This could be accomplished by driving
a2
p
transition on a second order red sideband, at
frequency
v
0
v
1
v
2
, on a particular (extra) ion “g”
which is initially in the ul
g
state. This ion is not used
to store information; it is only used for this one particu-
lar purpose. This would be followed by operations
which map the motional states back onto the internal
states of ions j and k (like step (2c) above). Overall, this
provides a phase gate (Eq. (54) with f =
p
) between
ions j and k. To make a CN gate between ions j and k,
we need to precede the above operations with a
p
/2
pulse on the internal state of ion j (or k) and follow the
above operations with a
p
/2 pulse on the internal state
of ion j (or k).
In this section, we have assumed that each ion can be
addressed independently. Also, since very many such
operations will be desired for a quantum computer, the
accuracy or fidelity of these operations is of crucial
importance. These issues are confronted in Sec. 4. As
noted in Sec. 2.3, in each separate operation involved in
a quantum computation, such as application of the red
sideband in step 2(a) to ion j, a definite phase of the
applied fields is assumed. This phase for each ion can be
chosen arbitrarily for the first operation, but upon
successive applications of the same operation to the
same ion, it must be held fixed, or at least be known,
relative to the initial phase. An exception to this is appli-
cation of 2
p
-pulses as in step 1(b) where the phase of
the fields does not enter into the final result of the
operation.
3.4 Entangled States for Spectroscopy
A collection of atoms, whose internal states are entan-
gled through the use of quantum logic, can improve the
quantum-limited signal-to-noise ratio in spectroscopy.
Compared to the factorization problem, this application
has the advantage of being useful with a relatively small
number of ions and logic operations. For example, for
high-accuracy, ion-based frequency standards [74,189,
190,191], use of a relatively small number of trapped
ions (L # 100) appears optimum. As outlined here, the
states involving L ions which are useful for spectroscopy
and frequency standards can be generated with L logic
gates. For L . 100, a significant improvement in perfor-
mance in atomic clocks could be expected.
In spectroscopy experiments on L atoms, in which the
signal relies on detecting changes in atomic populations,
we can view the problem in the following way using the
spin-1/2 analog for two-level atoms. The total angular
momentum of the system is given by J =
O
L
i=1
S
i
where S
i
is the spin of the ith atom. The basic task is to measure
v
0
, the frequency of transitions between the ul and ul
states Eq. (11). We first prepare an initial state for the
spins. We assume spectroscopy is performed by apply-
ing (classical) fields of frequency
v
R
for a time T
R
according to the method of separated fields by Ramsey
[173]. The same field is applied to all atoms. After
applying these fields, we measure the final state popula-
tions; that is, we detect, for example, the number of
atoms N
in the ul state. In the spin-1/2 analog, this
is equivalent to measuring the operator J
z
, since
N
= J
II
J
z
where
II
is the identity operator. We assume
the internal states can be detected with 100 % efficiency
(Sec. 2.2.1). If all sources of technical noise are elimi-
nated, the signal-to-noise ratio (for repeated measure-
ments) is fundamentally limited by the quantum fluctua-
tions in the number of atoms which are observed to be
in the ul state. These fluctuations can be called quan-
tum projection noise [100]. Spectroscopy is typically
performed on L initially nonentangled atoms (for exam-
ple,
C
(t =0)=P
L
i=1
ul
i
). With the application of the
Ramsey fields, the atoms remain nonentangled. For this
case, the imprecision in a determination of the fre-
quency of the transition is limited by projection noise to
(D
v
)
meas
= 1/(LT
R
t
)
1/2
where t >> T
R
is the total averag-
ing time [100]. If the atoms can be initially prepared in
particular entangled states, it is possible to achieve
(D
v
)
meas
< 1/(LT
R
t
)
1/2
. Initial theoretical investigations
for ions [3, 9] examined the use of correlated states
which could achieve (D
v
)
meas
< 1/(LT
R
t
)
1/2
when the
population (J
z
) was measured. These states are
analogous to those previously considered for interferom-
eters [192, 193]. More recent theoretical investigations
[194] consider the initial state to be one where, after the
first Ramsey pulse, the internal state is the maximally
entangled state
C
=
1
Ï2
(ul
1
ul
2
... ul
L
+e
i
f
(t)
ul
1
ul
2
...ul
L
),
(60)
where
f
(t)=
f
0
L
v
0
t. After applying the Ramsey
fields, we measure the operator O
ˆ
= P
L
i=1
S
zi
instead of
J
z
. This gives (D
v
)
meas
= 1/(L
2
T
R
t
)
1/2
, which is the max-
imum signal-to-noise ratio possible and corresponds to
the Heisenberg limit [194]. In the language of quantum
––
––
278
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
error correction, if we express
C
in terms of the basis
states u0l' =(u0l + u1l)/2
1/2
and u1l' =(u0l u1l)/2
1/2
,we
find that
f
(t) is determined from a parity check of the
total state in this second basis [142]. For an atomic clock
where the interrogation time T
R
is fixed by other con-
straints, this means that the time
t
required to reach a
certain measurement precision is reduced by a factor of
L relative to the nonentangled atom case. This improve-
ment is of significant practical importance since, to
achieve high measurement precision, atomic clocks are
run for averaging times
t
of weeks, months, and even
longer.
Cirac and Zoller [1] have outlined a scheme for pro-
ducing the state in Eq. (60) using quantum logic gates.
Starting with the state
C
(t =0)=P
L
i=1
ul
i
u0l,wefirst
apply a
p
/2 rotation (
V
0,0
t =
p
/4,
f
1
=+
p
/2 in Eq. (23))
to ion 1 to create the state
C
=2
–1/2
(ul
1
+ ul
1
)ul
2
ul
3
...ul
L
u0l. We then apply the CN gate C
ˆ
1,i
, sequentially
between ion 1 and ions i = 2 through L to achieve the
state of Eq. (60). An alternative method for generating
this state is described in Ref. [194]. As a final example,
we consider a method for generating the maximally
entangled state which requires a fixed number of steps,
independent of the number of ions. For simplicity, we
illustrate the method for three ions. Starting with the
state
C
(t =0)=ulululu0l, we first selectively drive
one of the ions (say ion 3) with a carrier
p
/2 pulse
followed by a red sideband
p
pulse to give the sequence
ulululu0l 2
–1/2
ulul(ul + ul) u0l 2
–1/2
ulul
ul(u0)l + u1l). We now use the Lamb-Dicke dependence
of the carrier transition (Eqs. (56)) to make an odd
(even) integer number of
p
flips correlated with the
un =0l state and an even (odd) integer number of
p
flips
correlated with the un =1l state (the laser beam intensity
is assumed to be the same on all ions). We can now
employ a transition to an auxiliary level. For example, if
the state after the last step, is the state 2
–1/2
( ulululu1l
+ ulululu0l), we could apply a blue sideband
p
pulse
between states ul and uauxl of one of the ions (say the
third) followed by a carrier
p
pulse on this transition to
carry out the steps 2
–1/2
(ulululu1l + ulululu0l)
2
–1/2
(ululuauxl + ululul)u0l) 2
–1/2
(ululul +
ululul)u0l. (If the state after the previous step was
2
–1/2
(ulululu0l + ulululu1l), we would sandwich
the last operation between
p
pulses on the ul ul
transition of the selected ion.)
Other correlated states can also be useful for spec-
troscopy. A strategy which essentially measures the
variance of N
˜
is discussed by Holland and Burnett
[195] and Kim et al. [196]. This method has also been
incorporated into a proposed technique for spectroscopy
of internal states of Bose-Einstein condensates [197].
In comparing the case for entangled vs non-entangled
states in spectroscopy, the above discussion has assumed
that T
R
is fixed. This constraint would be valid if the ions
were subject to a constant heating rate and we desired to
maintain the second-order-Doppler (time dilation) shift
below a certain value, for example. However, the use of
entangled states may not be advantageous, given other
conditions. For example, Huelga et al. [198] assume that
the ions are subject to a certain dephasing decoherence
rate (decoherence time less than the total observation
time). In this case, there is little advantage of using
maximally entangled states over non-entangled states.
The basic reason is that the maximally entangled state
decoheres L times faster than the states of individual
atoms. Therefore, when using the maximally entangled
state, T
R
must be reduced by a factor of L for optimum
performance. Because of this, the gain from use of the
maximally entangled state is offset by the reduced value
of T
R
. (Huelga et al. [198] actually show that a modest
improvement can be obtained under these conditions by
use of partially entangled states.) In appendix A, we
compare entangled vs nonentangled states in the context
of a practical atomic clock application where a refer-
ence oscillator is locked to the atomic resonance.
4. Decoherence
The atomic motional and internal states, and the
(logic) operations, were described above in an idealized
fashion. In this section we consider some of the practical
limitations to these idealizations. These limitations can
generally be grouped under the heading of decoherence
if, by decoherence, we mean any effect which limits the
fidelity of these operations (see Sec. 4.3). This is a more
general use of the term decoherence; in some treat-
ments, decoherence refers only to dephasing of qubit
states and does not include state changes. Although
somewhat arbitrary, we also find it convenient to break
decoherence into categories: (1) decoherence of the ion
motion, (2) decoherence of the ion internal levels, and
(3) decoherence caused by nonideal applied fields
which are responsible for the logic operations.
4.1 Motional Decoherence
For the trapped ion system discussed in this paper,
decoherence may be dominated by that of the motional
state. Scaling of decoherence will depend on the physi-
cal system being treated and the mechanism of decoher-
ence [199, 200]. For quantum computation with ions,
motional decoherence is somewhat easier to character-
ize than for a general motional state since we are
primarily interested in relaxation of the un =0l and
un =1l motional states for a particular mode (for exam-
ple, the center-of-mass (COM) mode along the axis of a
linear trap).
279
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.1.1 Phase Decoherence Caused by Unstable Trap
Parameters
A simple form of motional decoherence is caused by
fluctuations in trap parameters. Most likely, by employ-
ing electronic filtering, these parameters fluctuate
slowly on the time scale of the basic operations
(. 1/
V
n',n
), therefore, the motion is subject to dephasing
due to the corresponding adiabatic changes in motional
frequency. In the linear trap, if we assume a
i
<< q
2
i
,
i[{x, y} (Sec. 2.1), then for small fluctuations
d
V
0
,
d
U
0
,
dV
T
,
dk
,and
d
R,wehave
d
v
x,y
/
v
x,y
=
dV
0
/V
0
d
V
T
/
V
T
–2dR /R,
d
v
z
/
v
z
= 1/2(d
k
/
k
+ dU
0
/U
0
) . (61)
The relationship between these frequency fluctuations
and phase fluctuations in a series of logic operations is
discussed in Sec. 4.3.2. The effects of modulation of
these parameters by high-frequency noise is considered
in Sec. 4.1.3 and in Ref. [66]. Although an experimen-
tally open question, it is expected that all of these
parameters could be controlled sufficiently well that
they should not be the primary cause of decoherence.
4.1.2 Radiative Decoherence
Decoherence has received considerable attention in
connection with quantum measurement [201] and has
been put forth as a practical solution to the quantum
measurement problem [55,56,57]. In quantum optics, a
paradigm for decoherence has been to consider relax-
ation of the harmonic oscillator associated with a single
mode of the radiation field by coupling to the environ-
ment [202–206]. This kind of fundamental decoherence
has recently been observed in the context of cavity QED
by Brune et al. [207]. An important result from these
studies is that relaxation of superposition states occurs at
a rate which increases with the separation of the states
in Hilbert space and almost always precludes the exis-
tence of “large” Schro¨dinger-cat-like states except on
extremely short time scales.
A fundamental source of decoherence for the COM
mode of ion motion is understood by considering that
the dipole associated with the oscillating charged ion(s)
is radiatively coupled to the thermal fields of the envi-
ronment, at temperature T. The master equation describ-
ing the evolution of the density operator
r
for the motion
(in the interaction picture) can be written [208,209,210]
r
Ù =
g
2
(n¯ + 1) (2a
r
a
a
a
r
r
a
a)
+
g
2
n¯(2a
r
a aa
r
r
aa
) , (62)
where n¯ is the mean number of motional quanta when
the ion is in equilibrium with the environment
(n¯ = [exp(
"v
z
/k
B
T)–1]
–1
), and
g
is the relaxation rate of
the energy to thermal equilibrium. Since ion trap exper-
iments will typically operate in the situation where
"v
z
<< k
B
T, then n¯ . k
B
T/
"v
z
. We will assume the
ion(s) start in the subspace of density matrix elements
r
00
,
r
01
,
r
10
,and
r
11
where
r
ij
ki|
r
ujl. Equation (62)
implies
r
Ù
01
(t =0)= Ï2
g
(n¯+1)
r
12
g
(2n¯ + 1/2)
r
01
. –2n¯
gr
01
,
r
Ù
00
=–
g
[n¯
r
00
–(n¯+1)
r
11
],
r
Ù
11
=2
g
(n¯+1)
r
22
g
(3n¯+1)
r
11
+
g
n¯
r
00
. (63)
General expressions for d
r
/dt and dknˆl/dt are given in
Appendix B. Based on these expressions and in the limit
that n¯ >> 1, we will characterize the motional decoher-
ence by the time t* = 1/(n¯
g
), which is approximately the
time for the ion to make a transition from the ground
state. This agrees with a classical estimate [103].
4.1.3 Radiative Damping/Heating
The electric dipole associated with the an ion’s COM
oscillatory motion will couple to thermal (black body)
or ambient radiation in the environment. However, since
the wavelength corresponding to ion oscillation frequen-
cies will typically be much larger than the trap electrode
spacings, this coupling can be described by lumped-cir-
cuit models [102, 103]. In these models, we assume the
ion’s motion induces currents in the neighboring elec-
trodes; these currents, in turn, couple to the resistance of
the electrodes or circuit elements attached to the elec-
trodes. In experiments where these resistances are pur-
posely made high to maximize damping of the ion COM
motion, observed time constants agree with the model
[77,78,79,102,103,104]. In the two (single-ion) experi-
ments which have been able to achieve cooling to the
n = 0 motional state, the measured value of t* was about
0.15 s for a
198
Hg
+
ion [44] and about 1 ms for a single
9
Be
+
ion [17,21,45,47,131,132,211]. In these experi-
ments, t* was intended to be made as long as possible;
however, the observed values of t* were considerably
shorter than what we would calculate from the model,
as shown below.
The model considers the electric-circuit equivalent
shown in Fig. 6(a). Effectively, the electric-dipole
oscillator formed by the ion COM motion can be consid-
ered to be confined in a cavity formed by the trap
electrodes. A useful representation of this situation is to
model the COM motion (in one direction) as a series
inductive-capacitive (lc) circuit which is shunted by the
280
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
Fig. 6. Schematic diagrams of the lumped circuit equivalents for the trap electrodes and trapped ion(s). (a) The
left part of the diagram shows, schematically, the electrodes for a Paul trap with hyperbolic electrodes and a
collection of trapped ions. On the right is shown the corresponding lumped circuit equivalent; C
T
represents the
inter-electrode capacitance (the combined effects of the capacitances shown in the left part of the figure); r
represents the resistive losses in the electrodes and connecting wires, l
L
and c
L
represent the equivalent inductance
and capacitance for the COM mode of oscillation in the z (vertical) direction [103]. (b) A schematic diagram of
the endcaps electrode for the trap of Jefferts et al. [211] which was used in the NIST experiments (the ring is
not shown). Induced currents in the z (vertical) direction are assumed to follow a path indicated by shading; the
resistance in this path represents r in part (a) of the figure. (c) The rf potential between ring and endcaps electrode
(or between pairs of rods as indicated in Fig. 1) is typically generated with a resonant rf step-up transformer. The
resistance in this transformer can, in principle, couple to the ion motion as discussed in the text.
capacitance of the trap electrodes as shown [9, 103].
The resistance r is due to losses in the electrodes and
conductors which connect the electrodes. The Johnson
noise associated with this resistance can heat the ions.
The equivalent inductance of the ion COM motion is
given by l
L
. md
2
/L(
a
q)
2
where d is the characteristic
internal dimension of the ion trap electrodes, L the num-
ber of ions, and
a
is a geometrical factor on the order of
1 which can be computed [212]. For traps with hyper-
bolic electrodes, if we consider motion in the z direc-
tion, d =2z
0
(the separation of the endcap electrodes)
[67–70] and
a
. 0.8 [213,214,215]. For the trap used in
the NIST experiments, where z
0
= 130 mm, l
1
. 60 000
henries! The resistance r yields a time constant
l
L
/r =1/
g
. This implies [9]
281
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
t
*
=
1
g
n¯
=
"
Q
k
B
T
=
"v
z
l
L
k
B
Tr
=
4m
"v
z
q
2
S
E
(
v
z
)
, (64)
where Q =
v
z
l
L
/r is the quality factor of the ion oscilla-
tor. The last expression in Eq. (64) shows t* in terms of
the spectral density of electric field fluctuations at the
site of the ion which can be written S
E
=4k
B
Tr(
a
/d)
2
where 4k
B
Tr is the Johnson noise voltage associated
with the effective resistor r. The trap reported in the
NIST experiments had the endcaps made of a single
piece of molybdenum as shown schematically in Fig.
6(b) (ring electrode not shown). We assume the induced
currents flow in the endcaps electrode as indicated in the
shaded portion of Fig. 6(b), where
d
S
is the skin depth.
This seems to be a conservative estimate since currents
will also flow in the sides of the endcap electrodes and
will not be confined to the skin depth [216], thereby
reducing the effective value of r. Taking the resistivity
of molybdenum to be
r
(Mo) = 5.7 3 10
–6
V cm, w
T
=
125 mm, and x
T
= 1 mm, we find r . 2
r
x
T
/
d
S
w
T
= 0.0415
V. If we assume T . 300 K, t* . 4.6 s, considerably
longer than the observed value of t*(. 1 ms). An alter-
native model for dissipation of charges moving parallel
to a nearby surface [200] predicts a much larger value
of t*. Lamoreaux [217] has derived an expression which
agrees with Eq. (64), however he chooses a value of r
higher than what we calculate.
Faster heating will occur if T>>300 K. This can be
expected at the relatively high powers delivered to the
step-up transformer used to generate V
0
(. 1Winthe
NIST experiments), but this alone cannot explain the
difference between what the model predicts and the
observed heating rate. Conversely, if the trap can be
operated at cryogenic temperatures, this kind of heating
should be substantially reduced.
Heating can occur in the axial and radial directions
due to the interplay of a stray static field (e.g., from
patch potentials on the electrodes) and noise on U
0
or V
0
near one of the secular frequencies. Here, we explain
what appears to be one important case, a fluctuation of
U
0
in the presence of a stray static field along the z
direction of a linear trap. This and other cases are dis-
cussed more fully in Ref. [66].
In equilibrium, the force on an ion from a stray static
field E = E
s
zˆ is balanced by the field from the trap given
by Eq. (2). We have E
s
=2
k
qU
0
z
equil
, where z
equil
is the
equilibrium position of the ion (here, we assume
z
equil
= 0 in the ideal case). A fluctuation in U
0
therefore
causes a fluctuation in the electric field seen by the ion.
We can characterize the spectral density of these field
fluctuations as S
E
(
v
)=(E
s
/U
0
)
2
S
U
0
(
v
) where S
U
0
(
v
)is
the spectral density of potential fluctuations. From
Eq. (64), we have
t
*
=
4m
"v
z
q
2
S
U
0
(
v
z
)
F
U
0
E
s
G
2
. (65)
For a very small linear trap where k . (0.3 mm)
–2
,and
for m = 9 u (e.g.,
9
Be
+
)and
v
z
/2
p
= 10 MHz, we have
U
0
. 17 V. For E
s
. 100 V/m and S
U
0
=(1nV)
2
/Hz (the
Johnson noise voltage from a 60 V resistor at room
temperature), we have t* . 430 s. Since t* ~
v
z
d
2
,we
see there is a premium on having a relatively large trap
with large values of U
0
to keep
v
z
as large as possible.
In the above, we assumed that the ions couple to the
surroundings through the oscillating electric dipole due
to their COM motion. In situations where the extent of
the ion sample is small compared to the distance to the
electrodes, the induced currents result dominantly from
the COM mode; therefore radiative decoherence from
modes other than the COM mode can be substantially
suppressed [1, 103]. For example, for two trapped ions
aligned along the z axis, we would expect electric fields
from stray (fluctuating) potentials on one of the end
electrodes to cause an excitation force on the z stretch
mode which is suppressed by a factor equal to the ratio
of the ion spacing to trap dimensions compared to the
force on the COM mode.
Fluctuations in V
0
and U
0
for E
s
= 0 can also cause
heating of the ions. These sources are discussed in Ref.
[66]. Heating might be caused by parametric processes.
For example, heating could be induced if the trap pseu-
dopotential is modulated (coherently or by noise) at
twice the secular frequency. This problem has been
treated by Savard et al. [218] in the context of optical
dipole traps for neutral atoms (a kind of Paul trap for an
electron to which the atomic core is attached). For the
conditions of the NIST
9
Be
+
experiments, this kind of
heating was estimated to be too small to account for the
observed value of t
*
[66].
4.1.4 Injected Noise
Noise from various ancillary electronic devices might
be injected onto the electrodes; this additional electronic
noise could then heat the ions. Added electronic noise
can be modeled as a resistor r in Fig. 6 that has a
temperature much higher than the ambient temperature.
These sources of noise can be tested by injecting noise
at a level equal to or above the ambient noise level and
looking for a shortening of t*. For this test to be valid,
we must have a reliable means of sensing the noise at the
trap electrodes. This may be difficult to achieve in prac-
tice, since, in the experiments, it is usually desirable to
filter the electrodes from the rest of the environment at
the motional frequencies. This was the case in the NIST
experiments, where electronic filtering at the motional
frequencies precluded the direct observation of voltage
noise on the electrodes.
282
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.1.5 Motional Excitation From Trap RF Fields
The rf fields used for trapping in a Paul trap can lead
to excitation of the ion motion. We will consider four
types of effects in which the rf micromotion can, indi-
rectly, cause heating. For the first type of effect consid-
ered, we will analyze heating of the axial motion of a
single ion in a conventional spherical-quadrupole Paul
trap; the results can be generalized to other cases such
as the heating of radial modes in a linear trap. For a
spherical quadrupole trap, motion in the z (axial)
direction has the same form as Eq. (4). If we assume a
potential V
0
cos(
V
T
t) is applied between the ring and
endcap electrodes, we have
z(t) . A
z
cos(
v
z
t +
f
z
)
F
1+
q
z
2
cos(
V
T
t)
G
, (66)
where A
z
and
f
z
are set by initial conditions,
q
z
8qV
0
/(m
V
2
T
(r
2
0
+2z
2
0
)),
v
z
. q
z
V
T
/8
1/2
, r
0
is the inner
radius of the ring electrode, 2z
0
is the distance between
endcaps, and we assume q
z
<< 1. In the radial direction,
the motion will be similar with radial secular frequency
v
r
=
v
z
/2. From this equation, we see that the ion’s
motion in the z direction has components at frequencies
v
=
V
T
6
v
z
. Since the rf voltage V
0
is typically applied
through a resonant step-up transformer [shown sche-
matically in Fig. 6 (c)] the ion’s motion at these frequen-
cies might be expected to couple to the resistance R
T
between the ring and endcaps associated with this step-
up transformer. At a frequency
v
near
V
T
, the
impedance between the ring and endcaps electrode can
be represented by a parallel tuned circuit as shown in
Fig. 6(c). This impedance is given by
Z(
v
)=
R
T
1+2iQ
S
v
V
T
V
T
D
r
s
(
v
)+iX(
v
) , (67)
where Q is the quality factor for the circuit
(Q . R
T
/(
V
T
L
T
)=R
T
V
T
C
T
). Coupling to the effective
series resistance r
s
(
v
) should not occur if the endcap
electrodes are placed symmetrically around the ring
electrode (as intended). However, since the relative elec-
trode positions are difficult to control in small ion traps,
a displacement of the ring electrode toward one of the
endcaps will cause a net induced current from ion
motion to flow between ring and endcaps at frequencies
v
=
V
T
6
v
z
. We characterize this current (in the z
direction) by I =
b
qz˙ (2z
0
), where z˙ is the ion velocity
and
b
is a geometrical parameter which expresses the
coupling to the electrodes (
b
= 0 when the ring is placed
symmetrically between endcaps). The effective induc-
tance of the ions for this type of coupling is given by
l'
L
= l
L
/
b
2
. Associated with r
s
(
v
) in a small bandwidth
D
n
around
v
is a series Johnson noise kV
2
n
l =4k
B
TD
n
r
s
.
The electric field associated with this noise at frequen-
cies
v
=
V
T
6
v
z
can heat the ion motion in a way
similar to the way in which the motion can be excited by
a coherent excitation at these frequencies [68]. From
Eq. (3.7) of Ref. [68], we see that an electric field E
1
zˆ
applied at a frequency
V
T
6
v
z
is equivalent to an
electric field [
v
z
/(
V
T
Ï2)]E
1
zˆ applied at frequency
v
z
.
Therefore, the Johnson noise from the series resistance
r
s
(
v
=
V
T
6
v
z
) is equivalent to that from a series
resistance r'
s
=[
v
2
z
/(2
V
2
T
)]r
s
(
v
=
V
T
6
v
z
) at frequency
v
z
. The heating from this source is characterized by the
heating time t*' = l'
L
/(r'
s
n¯). For the NIST single
9
Be
+
ion
experiments, this source of heating was estimated to be
negligible.
A second type of rf heating can occur due to the
Coulomb interaction between ions. In a collection of
ions, such a string of ions in an ideal linear trap, the
Coulomb coupling between ions makes all of the
motional modes, except the COM modes, nonlinear.
This can lead to excitation of these modes in a Paul trap
by the driving fields at frequency
V
T
. This excitation
and the resulting chaotic motion have been studied
extensively for two ions trapped in a conventional Paul
trap. Experiments have been performed at Munich and
IBM; these studies are discussed in Ref. [219]. More-
over, even for a single (harmonically bound) ion, nonlin-
ear subharmonic excitation can occur if the exciting
field is inhomogeneous [220]. Both types of heating can
be made negligible when the mode frequencies are not
submultiples of
V
T
, when a
i
, q
2
i
<< 1, and when all
modes are sufficiently cooled and therefore very linear.
Another type of rf heating occurs in some experi-
ments when the condition a
i
, q
2
i
<< 1 is not rigorously
satisfied and the trapped ions are fairly energetic. The
motion of single ions (or multiple ions when the mutual
Coulomb interaction or “space charge,” can be
neglected) will be unstable when the condition
p
v
z
+ m
v
r
=
V
T
is satisfied in a spherical quadrupole
trap or when p
v
x
+ m
v
y
=
V
T
in a linear trap) (p and m
are integers). This type of heating has been observed in
some beautiful experiments [221] and has been ex-
plained theoretically [222]. These “heating resonances”
arise from terms in the trap potential which are higher
order than quadratic. We briefly explain their origin for
a single ion. For simplicity, we neglect the contribution
of the static potentials; their inclusion will not change
the analysis significantly.
283
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
In general, the potential of the trap can be expanded
about the equilibrium position of the ion and written in
spherical coordinates (r,
u
,
f
)as
F
=V
0
cos
V
T
t
O
`
l=0
O
l
m=–l
C
l,m
F
r
d
G
l
Y
l,m
(
u
,
f
) , (68)
where Y
l,m
are the spherical harmonics and d is a charac-
teristic dimension of the trap. We take d = R for the
linear trap (and d =(r
2
0
+2z
2
0
)
1/2
for the spherical quadru-
pole trap) [67,68,69,70]. For the ideal linear trap
(Eq. (1) with U
r
= 0, and neglecting a constant term),
only two terms in the expansion in Eq. (68) contribute,
C
2,2
= C
2,–2
=–(2
p
/15)
1/2
. (For the ideal spherical
quadrupole trap, neglecting a constant term, only one
term contributes, C
2,0
=4(
p
/5)
1/2
.) For a nonideal linear
trap, the resonance heating can be explained as due to
terms in Eq. (68) which give rise to a Hamiltonian of the
form
H
I
= q
F
= qV
p,m
F
x
R
G
p
F
y
R
G
m
cos
V
T
t . (69)
In the interaction picture for the motion, this becomes
H'
I
.
qV
p,m
2
F
x
0
R
G
p
F
y
0
R
G
m
S
(a
x
)
p
(a
y
)
m
+ h.c.
D
(70)
where this last expression is the leading term which
satisfies the resonance condition p
v
x
+ m
v
y
=
V
T
. This
interaction will be suppressed because of the inherent
smallness of high-order anharmonic terms V
p,m
(for sim-
ple trap electrode shapes) and the smallness of the terms
(x
0
/R)
p
and (y
0
/R)
m
. Furthermore, if the xymodes are
cooled to near the zero-point energy, matrix elements of
the motional operators will be near 1. For large ampli-
tudes of motion, the mode frequencies are not well de-
fined because of anharmonic terms and heating from
this coupling would be expected to occur. In any case,
it is easy to check for a resonant heating of this type by
varying the resonant frequencies
v
x
and
v
y
relative to
V
T
. It can also be checked by varying the initial ampli-
tudes of motion in the modes. These tests were used for
the NIST single
9
Be
+
ion experiments; no change in the
heating was observed.
Similarly, in a linear trap with L ions, we would
expect resonances to occur when
O
3L
k =1
m
k
v
k
=
V
T
where m
k
are integers and
v
k
are normal mode frequencies (see
Sec. 2.3.2). If all modes are cooled to the point where
only a few motional states are excited, then the mode
absorption spectrum will consist of sharp features
around the mode frequencies, and the resonances can be
avoided by changing the trap parameters. Moreover, the
coupling parameters (Sec. 4.1.8) will, in general, be
very small.
A fourth type of heating due to rf trap fields is ex-
plained as follows. A common problem in ion trap ex-
periments is the presence of stray static electric fields.
These fields can give rise to coherent motion at fre-
quency
V
T
and potential to heating, which must be ac-
counted for. Stray static electric fields can arise from
potential variations on the electrode surfaces (“patch”
fields) due, for example, to the finite crystalline grain
size of the electrode material [223], or charge buildup
on the trap electrodes. Charge buildup can occur be-
cause, typically, ions are created by electron impact
ionization of neutral atoms which pass through the trap.
Often, the ionizing electrons are also collected by the
electrode surfaces. Electrode charging is particularly
important at low temperatures where, apparently, ad-
sorbed gases on the electrodes can provide an insulating
surface upon which stray charge resides for long periods
of time (hours).
If stray static electric fields are present, the equi-
librium position of an ion is shifted to a place where the
force from the stray field is counterbalanced by the force
from the pseudopotential. We will analyze the effects of
such stray fields using a classical treatment of the
motion of a single trapped ion. In general, E
stray
= E
sx
xˆ + E
sy
yˆ + E
sy
zˆ. Stray fields along the z direction in a
linear trap merely shift the origin along this direction
and can therefore be neglected. For balance in the x and
y directions, we have F
x
= qE
sx
(q
F
p
)/
x =0and
F
y
= qE
sy
(q
F
p
)y = 0 where
F
p
is given by Eq. (6).
This leads to offsets Dx and Dy and resultant motions
which are, to first order in q
x
and q
y
given by
x(t) . (Dx + A
x
cos(
v
x
t +
f
x
))
F
1+
q
x
2
cos(
V
T
t)
G
,
Dx =
qE
sx
m
v
2
x
,
y(t) . (Dy + A
y
cos(
v
y
t +
f
y
))
F
1+
q
y
2
cos(
V
T
t)
G
,
Dy =
qE
sy
m
v
2
y
, (71)
where, as in Eq. (4), A
x
and A
y
are the amplitudes of
secular motion. The presence of the offsets Dx and Dy
means that the ion motion has an additional component
at frequency
V
T
. This motion will effectively give rise
to sidebands on the applied radiation as seen by the ion,
thereby reducing the size of matrix elements between
states. For example, for single-photon transitions driven
by a traveling wave with wavevector k = k
x
xˆ + k
y
yˆ , the
electric field from this traveling wave at the site of the
284
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
ion is proportional to exp(i[k
x
x(t)+k
y
y(t)–
v
t +
f
]) +
c.c. Terms in the exponential like k
x
Dx are (constant)
phase shifts which can be neglected. Terms like A
x-
cos(
v
t +
f
x
)[1+(q
x
/2)cos(
V
T
t)] are just the motion of
the ions in the ponderomotive potential in the ideal case.
The factor from the remaining term,
exp(i
q
x
2
[k
x
Dx k
y
Dy] cos(
V
T
t))
= cos[
f
V
cos(
V
T
t)] + isin[
f
V
cos(
V
T
t)] , (72)
where
f
V
(q
x
/2)[k
x
Dx k
y
Dy] gives rise to frequency
modulation sidebands on the spectrum which are spaced
by
V
T
. If we consider the carrier or central part of this
spectrum (the first term in the expansion of
cos[
f
V
cos(
V
T
t], we find that the matrix elements are
reduced by the factor J
0
(
f
V
) compared to the case
where the static fields are absent. For
f
V
<< 1, J
0
(
f
V
)
. 1–(
f
V
/2)
2
. exp(– (
f
V
/2)
2
). Therefore the effect of
the micromotion looks like an additional Debye-Waller
factor due to the smearing out of the atom’s position
over the exciting wave (see the discussion following Eq.
(25) and Sec. 4.3.5).
To the extent that offset fields are constant, they
should not cause heating unless the
V
T
sidebands give
rise to unwanted spectral components that are close to
transition frequencies of interest. However, the offsets
Dx and Dy can lead to a problem if the trapping field has
noise V
n
at frequencies
V
T
6
v
r
, that is, V
0
cos
V
T
t
V
0
cos
V
T
t + V
n
cos(
V
T
+
v
r
)t + V
n
cos(
V
T
v
r
)t. In this
case (assuming Dy = 0 for simplicity), the ion experi-
ences noise fields at frequencies
V
T
6
v
r
equal to
E
x
=–
F
/
x =–V
n
Dx/R
2
[Eq. (1)]. From the first part
of this section, this is equivalent to noise fields of ampli-
tude V
n
Dxq
x
/(4R
2
) applied at frequency
v
r
[66]. For the
NIST single
9
Be
+
ion experiments, this effect was esti-
mated to be negligible. Moreover, this source of heating
was tested for by purposely applying a static field offset
and seeing if the observed heating rate increased; a null
result was obtained. Experimentally, it has been possible
to reduce stray static electric fields by heating the elec-
trodes [191, 224] or cleaning the electrodes with elec-
tron bombardment [225]. Alternatively, they can be
compensated for with the use of correction electrodes
[226–228].
4.1.6 Fluctuating Patch Fields
Electrode patch fields might also vary in time; if the
spectrum of these variations overlaps the mode frequen-
cies, this could lead to ion heating. Investigations into
patch fields have primarily been done for time indepen-
dent or very slowly varying (< 500 Hz) components
[223]. However, fluctuating patch fields caused by fluc-
tuating adsorbate coverage has been studied in some
cases [229, 230]. These studies differ somewhat in the
low frequency behavior at time scales comparable to
diffusion times, but at frequencies
n
>
n
c
, where 1/
n
c
is
a time constant characteristic of surface diffusion, they
predict S(
F
n
,
n
) ~
n
a
where S(
F
n
,
n
) is the spectral
density of rms potential fluctuations
F
n
(in units of
V
2
–Hz
–1
)and
a
. 3/2 [230]. To estimate the effects
of time-varying patch potentials on a single trapped ion,
we assume the ion is sensitive only to the potential on a
portion of a nearby electrode. We take the area of this
portion equal to
p
a
2
p
where a
p
is the distance between
the ion and the nearest part of the electrode surface. The
effects of these potentials on the ion motion in one
direction is then estimated by assuming the ion is cen-
tered between two capacitor plates of area
p
a
2
p
sepa-
rated by a distance 2a
p
. The fluctuating potentials on
these plates give rise to a fluctuating field at the site of
the ion which can then excite its motion.
Patch potential fluctuations can be caused by the
fluctuations in the surface coverage from adsorbed
background gas molecules (or atoms). High frequency
fluctuations appear to be dominated by surface diffu-
sion rather than adsorption and desorption [230]. The
number of adsorbed molecules in an area
p
a
2
p
can be
approximated by N(
u
) .
u
(
p
a
2
p
)/(
p
r
2
a
) where
u
is the
fractional coverage and r
a
is the radius of the adsorbed
molecule. For low coverages (
u
<< 1), the number of
molecules will fluctuate randomly, DN . N
1/2
, which
leads to fluctuations in coverage D
u
.
u
1/2
r
a
/a
p
. A sim-
ple model for changes in the surface potential due to
adsorbed molecules is that the molecules are polarized
by the surface and effectively screen the surface poten-
tial. We can relate the change in potential of the plate to
the change in surface coverage by D
F
. kD
u
= k
u
1/2
r
a
/
a
p
, where k is a proportionality constant.
If we take S(
F
n
,
n
)=S
0
(constant) for
n
<
n
c
and
S(
F
n
,
n
)=S
0
(
n
c
/
n
)
3/2
for
n
>
n
c
,wehavek
F
2
n
l = eS(
F
n
,
n
)d
n
=3S
0
n
c
. Here k
F
2
n
l
1/2
is taken to be equal to the
value of D
F
estimated in the previous paragraph. The
cutoff frequency
n
c
is given by
n
c
. 1/t
diff
where t
diff
can
be approximated by t
diff
= l
2
d
/(4D) where D is the diffu-
sion constant and l
d
is the diffusion length [231]. Here,
we take l
d
to be the radius of the effective patch (l
d
. a
p
).
We then find for
n
>
n
c
,
S(
F
n
,
n
)=4
Q
D
1/2
F
(
k
r
a
)
2
3a
3
p
G
n
–3/2
, (73)
where an extra factor of 2 has been included to account
for the two capacitor plates which are placed on either
285
Volume 103, Number 2, March–April 1998
Journal of Research of the National Institute of Standards and Technology
side of the ion. To calculate the heating rate from these
potential fluctuations, we first note that they will pri-
marily act over a narrow bandwidth associated with the
ion’s motional frequency. In this case, we can represent
the fluctuations as coming from the Johnson noise of a
resistor r at temperature T connected between the capac-
itor plates, that is, S(
F
n
,
n
)=4k
B
Tr (assuming the
capacitive impedance is much greater than r). There-
fore, we can rewrite Eq. (64) as
t*=
4
"v
z
l
L
S(
F
n
,
n
)
, (74)
where
v
z
is the ion oscillation frequency.
Since the polarizability of molecules and atoms does
not change dramatically for different species, we will
estimate k from a measurement of change in surface
potential for potassium atoms on tungsten. From Fig. 2
of Schmidt and Gomer [232], we find k . 3 V. For an
estimate of
u
, we extrapolate the data presented in
Fig. 6.6 of Tompkins [233] for H
2
on tungsten and find
u
. 0.13 at a partial pressure of 10
–8
Pa. The diffusion
constant for H
2
on Mo is approximately equal to 10
–11
cm
2
? s
–1
[233]. To make a comparison with the heating
observed on a single
9
Be
+
ion in the NIST experiments,
we take l
d
. a
p
. 130 mm,
v
z
/2
p
= 11 MHz, L =1,
l
1
= 6.23 10
4
H, r
a
= 1 nm (Sec. 4.1.3), and we find
n
c
. 2.4 3 10
–7
Hz and t* . 3000 s.
This model is very sensitive to the high frequency
dependence of S(
F
n
,
n
)on
n
, and because of the very
low value of
n
c
estimated here, the model should be
refined. However, we note that if the value of t*(. 1
ms) for
9
Be
+
observed in the NIST experiments is caused
by fluctuating potentials on the surfaces of the elec-
trodes, this would correspond to S(
F
n
, 11 MHz)
. (1.3 nV)
2
/Hz. This should be detectable with a sensi-
tive amplifier. Therefore, independent of the model, this
type of noise may be detectable in a straightforward way.
Conversely, we note that a single trapped ion in the
experiments considered here will be an extremely
sensitive detector of potential fluctuations on electrodes
in vacuum.
4.1.7 Field Emission
Field emission from sharp protrusions on the elec-
trode surfaces can cause ion heating, either from the
direct electron-ion Coulomb coupling or from associ-
ated electronic noise on the electrodes. Field emission
caused by the trap potentials is not unexpected, and field
emission points have been observed to grow in a number
of ion trap experiments. For typical values of V
0
and trap
dimensions, the electron transit times from one
electrode surface to another are much less than 1/
V
T
so
that field emission occurs as if the fields are quasi-static.
Since the onset of field emission varies exponentially
with the applied voltage between electrodes, it is
possible to check for field emission by varying the trap
potentials by small factors, and monitoring the ion heat-
ing rate. This technique appears to have ruled out field-
emission heating in the NIST experiments since the
change in ion heating was much less than a factor of two
when V
0
was reduced by a factor of two. This argument
assumes the exponential variation of field emission with
applied voltage; if the emission or some leakage current
is less sensitive to voltage changes, this test may not be
valid. If field emission points are formed, it is usually
straightforward to remove them by momentarily apply-
ing a large negative potential to the electrode in
question. The resulting high current is usually sufficient
to “burn out” the field emission tip.
4.1.8 Mode Cross-Coupling From Static Electric
Field Imperfections
According to the scheme of Cirac and Zoller [1], the
operations which provide quantum entanglement of the
internal states of L trapped ions involve the coherent
manipulation of a single mode k of collective motion. In
the quantum logic scheme discussed in Sec. 3.3, this
mode is typically taken to be the COM mode along the
axis of a linear trap. A potential source of motional
decoherence is caused by the coupling of this kth mode
to one or more of the 3L-1 other spectator modes of
vibration in the trap. If the 3L-1 other modes of oscilla-
tion are not all laser-cooled to their zero-point energy,
then energy can be transferred to the kth mode of inter-
est. Even when the spectator modes are cooled to the
zero-point state, they can act as a reservoir for energy
from the COM mode. This does not lead to heating but
can cause decoherence. Ideally, the ions are subjected to
quadratic potentials as in Sec. 2.1. In practice, higher-
order static potential terms are present; these terms can
induce a coupling between the modes. Similar couplings
are induced by the intentionally-applied time-varying
fields necessary for providing entanglement; these are
discussed in Sec. 4.4.7 below.
We will assume that the higher order field gradients
act as a perturbation to the (harmonic) normal mode
solution. Following the convention of Eq. (30), these
fields will be specified by E
i
for i [{1,2,...3L} where
the index i specifies both the ion and direction of E.We
write the electric field at the jth ion as
E
j
= E
j
xˆ+E
L+j
yˆ + E
2L+j
zˆ , j[{1,2,. . . L} . (75)
286
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
From Eq. (31), we can write the equation of the kth
normal mode as [109, 234]
2
q
k
t
2
+
v
2
k
q
k
=
q
m
O
3L
i=1
D
i
k
E
i
. (76)
In general, we can write
E
i
= E
i
({u
p
}) = E
i
({q
j
})
= E
i
({q
j
}=0)+
O
3L
m=1
q
m
F
E
i
q
m
G
{q
j
}=0
+
1
2
O
3L
l=1
O
3L
m=1
q
l
q
m
F
2
E
i
q
l
q
m
G
{q
j
}=0
+ . . . (77)
where the derivatives are evaluated at the equilibrium
positions. The first term on the right side of this equa-
tion just gives rise to a shift of the equilibrium positions,
and the second term can be absorbed into new defini-
tions of the normal mode frequencies
v
i
. The second-
order term (last term shown in this equation) can reso-
nantly couple two modes of oscillation (l and m)tothe
normal mode of interest k. We find a possible resonant
term:
2
q
k
t
2
+
v
2
k
q
k
=
q
m
O
3L
i=1
D
i
k
q
i
q
m
F
2
E
i
q
l
q
m
G
{q
j
}=0
,
(78)
where the l and m mode frequencies satisfy u
v
l
6
v
m
u =
v
k
. This type of coupling can either add to or extract
energy from mode k, depending on the relative phases
of motion in the three modes. By substituting the free
solution to modes l and m [q
j
(t)=Q
j
exp(6i(
v
j
t +
f
j
)]
into the last equation, we find that if q
k
(t =0)=
(dq
k
/dt)
t =0
= 0, the driven solution to the amplitude of
mode k initially grows linearly with time:
uq
k
(t)u =
)
qt
2m
v
k
O
3L
i=1
D
i
k
Q
l
Q
m
F
2
E
i
q
l
q
m
G
{q
j
}=0
)
.
(79)
We illustrate with an approximate numerical example
which might have been expected to play a role in the
heating that was observed in the NIST experiments.
In those experiments, performed on single
9
Be
+
ions, the
heating that was observed was such that the ion made a
transition from the n =0ton = 1 level in about 1 ms.
For a single ion, the three normal modes are just the
oscillation modes along the x, y,andz directions
(q
1
= x, q
2
= y, q
3
= z; D
i
k
=
d
i,k
). The mode frequencies
were (
v
x
,
v
y
,
v
z
)/2
p
. (11.2, 18.2, 29.8) MHz, thus
approximately satisfying the condition
v
x
+
v
y
=
v
z
. For
sake of argument, we assume this resonance condition to
be exactly satisfied. We consider heating of the x motion
assuming both the y and z modes are excited. From Eq.
(79), we find ux(t)u = uqtA
y
A
z
[
2
E
x
/yz]
(y = z =0)
/(2m
v
x
)u
where A
y
and A
z
are the amplitudes of motion in the y
and z directions. For simplicity, we neglect the fact that
energy is exchanged between the three modes; that is,
we assume the amplitudes of the y and z motion remain
fixed. In this approximation, if A
y
= A
z
=
j
, the time it
takes to excite the x motion to the same amplitude (
j
)
is given by t = u2m
v
x
/(q
j
[
2
E
x
/yz]
(y = z =0)
)u.If
j
.
10 nm (corresponding to knˆ
y
l.knˆ
z
l.1 for the condi-
tions of the single
9
Be
+
ion NIST experiments, the field
gradient required to drive the x motion to an amplitude
of 10 nm (knˆ
x
l.1) in the observed time of 1 ms is
approximately
2
E
x
/yz = 1000 V/mm
3
.Itishighly
unlikely the gradient was this large for the NIST exper-
iments, and, furthermore, the resonance condition was
only approximately satisfied. Moreover, this source of
heating was easily tested by varying the initial values of
A
y
and A
z
(by varying the Doppler-cooling minimum
temperature through laser detuning) and studying the
heating rate of the x motion which had previously been
cooled to the zero point of motion. No dependence on
the initial values of A
y
and A
z
was found. In any case, if
all modes of motion are initially cooled to the zero-point
state this source of heating vanishes because the as-
sumed coupling only provides an exchange of energy
between modes. This places a premium on cooling all
modes to as low an energy as possible. Finally, it appears
that this single-ion example gives a worst case scenario
since, for large numbers of ions, the force on the gener-
alized coordinates [right hand side of Eq. (76)] requires
a high-order field gradient to be nonzero. These gradi-
ents are highly suppressed in the typical case where
ion-ion separation is much smaller than the distance
between the ions and the trap electrodes.
4.1.9 Collisions With Background Gas
Although trapped ion/quantum logic experiments will
typically be carried out in a high-vacuum environment
(P <10
–8
Pa), residual background gas collisions can be
important. The effects of collisions can be broken up
into two classes: (i) inelastic collisions, which alter the
internal state of the trapped ion or even change the
species of the ion, and (ii) elastic collisions, which only
add kinetic energy to the ion. Both types of collisions
will cause decoherence, although heating from elastic
collisions is expected to be to be the chief concern.
287
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
The most troublesome inelastic processes are chemi-
cal reactions and charge exchange. A background gas
atom or molecule can collide and chemically react with
the trapped ion, creating a different ionic species which
is no longer useful. For the reactions to occur, they must
be energetically favorable (exothermic), and, if the back-
ground neutral is a molecule, the reactions almost always
proceed since the internal degrees of freedom of the
molecule can help satisfy energy and momentum con-
servation in the reaction. In the ion trap experiments
discussed here, the ion can spend an appreciable amount
of time in the (optically) excited state due to laser exci-
tation; in this case the extra energy due to laser excita-
tion can make an otherwise endothermic reaction be-
come exothermic. For example, in experiments on
laser-cooled Hg
+
ions [235], when ions were excited to
the metastable 5d
9
6s
22
D
5/2
level (approximately 4.4 eV
above the ground state) they reacted with neutral Hg
atoms in the background gas to cause loss of the Hg
+
ions (presumably due to radiative association causing
Hg
+
2
dimer formation). As a second example, in experi-
ments on
9
Be
+
ions [47,76,211], the ions were converted
to BeH
+
upon collision with an H
2
molecule when reso-
nant light was applied to the
2
S
1/2
2
P
1/2,3/2
transitions.
As a final example, the formation of YbH
+
by a similar
process has been carefully studied by Sugiyama and
Yoda [236]. The second form of inelastic collision is
charge exchange, where a neutral background atom
gives up an electron and neutralizes the trapped ion.
Both types of inelastic collision depend critically on the
particular constituents involved.
Chemical reactions and charge exchange can occur
only if the interparticle spacing of the two colliding
partners approaches atomic dimensions. An upper limit
on these rates is given by the Langevin rate, for which
background neutrals penetrate the angular momentum
barrier and undergo a spiraling-type collision into the
ion [237]. In these collisions, the electric field from the
trapped ion polarizes the background neutral (polariz-
ability a), resulting in an attractive interaction potential
U(r)=2
a
q
2
/(8p
e
0
r
4
). Impact parameters less than a
critical value b
crit
=(
a
q
2
/p
e
0
m
v
2
)
1/4
will result in spiral-
ing collisions, where
m
and v are the reduced mass and
relative velocity of the pair. (Since the ions are assumed
to be nearly at rest from laser cooling, v is simply given
by the velocity of the background constituent.) The ve-
locity-independent Langevin rate constant
k
Langevin
s
v = pb
2
crit
v leads to an overall reaction rate
g
Langevin
= nk
Langevin
= nqÎ
p
a
e
0
m
, (80)
where n is the density of the background gas. In a
metal/glass room-temperature apparatus such as was
used in the NIST
9
Be
+
experiments [47,76,211], the
dominant background gas constituent is usually H
2
.
For H
2
, we obtain k
Langevin
= 1.64 3 10
29
cm
3
2 s
21
.Ata
pressure of 10
28
Pa and a temperature of 300 K, we
have
g
Langevin
. 0.004 s
21
. Other candidate background
molecules and atoms have similar polarizabilities and
values of k
Langevin
. Experimentally, we observe lifetimes
of several hours for trapped
9
Be
+
and
24
Mg
+
ions at
pressures of around 10
28
Pa, indicating that at least for
ground state
9
Be
+
and
24
Mg
+
ions the probability of
chemical reactions with the background gas constituents
is small. In a cryogenic ion trap for
199
Hg
+
, the lifetime
is many days [74].
Background gas can heat the trapped ions by transfer-
ring energy during an elastic collision. The Langevin
rate above gives too low an estimate of the rate of colli-
sions which transfer energy to the ion, since a “heating”
collision need not penetrate the angular momentum bar-
rier. A conservative estimate for the heating rate can be
given from the more general expression for the total
collision cross section
s
elastic
(in the quasi-classical limit)
in a C
4
/r
4
potential [238]. We take C
4
=
a
q
2
/8p
e
0
and
find
s
elastic
= p
G
(1/3)
F
a
q
2
16
e
0
"
v
G
2/3
. (81)
If we average over a thermal distribution of background
H
2
velocities, this results in the rate constant k
elastic
=
<
s
elastic
v > = 1.23 3 10
5
a
2/3
(v˜)
1/3
where v˜ (2k
B
T/
m
)
1/2
.
For H
2
at 300 K and a pressure of 10
28
Pa,wefind
k
elastic
. 1.24 3 10
28
cm
3
2 s
21
and
g
elastic
. 0.03 s
21
.
Although each collision on average transfers a large
amount of energy to the trapped ion, we conclude that at
typical UHV pressures, such collisions will also be rare.
Collisional heating can be tested by raising the back-
ground gas pressure. A simple way to do this is simply
turn off the vacuum pump; collisions as a source of
heating in the NIST experiments was eliminated in this
way. However, doing so does not insure that all partial
pressures increase by the same factor. Preferably, the
partial pressure of selected gases should be increased by
leaking them into the vacuum system and looking for an
increase in ion heating.
When ions are first loaded into a trap, their kinetic
energy is in general comparable to the depth of the trap
(typically greater than 1 eV). In this regime, elastic
collisions with the background gas are actually benefi-
cial, as the background gas can provide a viscous damp-
ing medium and bring the temperature of the trapped
ions into thermal equilibrium with the surrounding gas
[239]. This allows initial laser-cooling to proceed much
faster.
288
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.1.10 Experimental Studies of Heating
Some experimental diagnostics for heating have been
discussed above in Secs. 4.1.4–4.1.9. As discussed in
Sec. 4.1.2, we are primarily interested in determining
the heating from the n =0l or n =1l states. In the
NIST single
9
Be
+
ion experiments, heating could be esti-
mated by first preparing the ion in the ↓l0l state,
waiting a certain delay time, and then measuring the
strength and ratio of the first red and blue sidebands
[44,45]. This method gives a simple indication of the
heating rate, but a more complete method is described
here.
More recently, we have determined that depletion of
the n =0l state was dominated by a nearly continuous
and smooth heating which initially took the atom from
n =0 to n = 1. This was established in experiments
where the distribution of n level populations was mea-
sured after a delay time using the technique of Ref. [21]
and the tomographic technique of Leibfried et al.
[131,132]. For example, in the experiments of Ref. [21],
many measurements were repeated for each value of the
time delay in order to extract the population distribu-
tions from P
(t) [Eq. (42)]. In this way the time t
*
[Eq.
(64)] was determined to be about 1 ms. In this same
apparatus, if we make the assumption that the heating
was caused by a coupling to the environment at 300 K,
the overall time constant to reach equilibrium is
t
equil
. t
*
(k
B
T/
"v
x
) . 570 s.
Interestingly, under typical operating conditions,
these tests of heating are highly insensitive to heating
caused by collisions. In the previous section, we saw that
at the typical background pressure of 10
28
Pa for the
experiments of the NIST single
9
Be
+
ion experiments,
collisions would be expected to occur at a maximum rate
of about 0.03 s
21
, as estimated from total elastic scatter-
ing. Strong heating collisions, given by the Langevin
rate, occur at a rate of about 0.004 s
21
. These rates
indicate a time constant to reach equilibrium with the
300 K background gas of between 30 s and 250 s,
shorter than the 570 s time estimated from the observed
n =0n = 1 heating rates (previous paragraph). How-
ever, since each heating-rate experiment takes only
about 1 ms, at most, only one experiment in about
30 000 indicates a background gas collision; these events
are simply lost in the experimental noise. In effect, the
technique of measuring t
*
is able to detect the continu-
ous, smooth heating occurring in between collisions,
even though collisions are expected to be a stronger
overall heating source. This effect was even more pro-
nounced in the experiments of Ref. [44].
Although the source of heating in the NIST experi-
ments is not understood at the present time, it should be
possible to tell if the heating is radiative or from some
other cause by comparing the heating rates of the axial-
COM and axial-stretch modes of two ions in the trap.
The heating of the COM mode should be nearly the
same as for a single ion since it can be excited by a
spatially uniform (oscillating) field. However, since the
stretch mode will be excited only by a field gradient,
radiative heating of this mode by fields emanating from
the trap electrodes should be significantly less.
4.1.11 Experimental Studies of Motional
Decoherence
It is desirable to have some methods to test for deco-
herence. Full characterization of decoherence in a mo-
tional state could be accomplished by reconstruction of
the density matrix [131,132] coupled with a time delay
between creation and measurement. This complete char-
acterization may not be necessary and other methods
have been used. One possibility is to create an interfer-
ence signal between two states that depends on decoher-
ence mechanisms and monitor the contrast of that signal
in time. This type of measurement was used to charac-
terize the purity of Schro¨dinger-cat states for trapped
atoms [47]. Decoherence of Schro¨dinger-cat states of
the electromagnetic field, caused by radiative damping,
has been monitored by studying a correlation between
observed states of two successive atoms which probe the
field [207]. One kind of Schro¨dinger-cat for motional
states has the form (
a
l + 2
a
l) where
a
l denotes a
coherent state. A measurement of the value of the
Wigner function at the origin W(
a
= 0) may be suffi-
cient to characterize decoherence [131,240]. If
a
is suf-
ficiently big (k0
a
l 0), we would expect an initial
nonzero value for W(0) that is damped toward zero in
the course of decoherence [240,241]. For quantum logic
with ions, it appears that decoherence of the COM mo-
tional states in the submanifold of states 0l and 1l is
of primary importance. It may therefore be sufficient to
reconstruct the 2 3 2 matrix spanned by the 0l and 1l
number states of the COM mode and characterize its
evolution due to decoherence. One scheme is outlined
here.
Suppose we initially create the state
C
(t =0)=
↓l(C
,0
(0)0l + C
,1
(0)1l). In a particular experi-
ment, after a time t
d
the state becomes
C
(t
d
)=
↓l(S
`
i=0
C
,i
(t
d
)il) due to motional decoherence. We
can then apply two analysis pulses of radiation to this
state. We first apply a p pulse on the ↓l1l →l0l
red sideband transition. We assume this pulse takes time
t
1
(
V
0,1
t
1
= p/2). We follow this with a p/2 pulse on the
↓l →l carrier transition. Here, we simplify the dis-
cussion by assuming the time for both of these pulses is
short enough that decoherence during the pulses can be
neglected. The wavefunction is now given by
289
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
C
final
=
1
Ï2
S
↓l
c
(x)+↑l
c
(x)
D
,
c
(x)=[C
,0
(t
d
) 2 e
i(
f
1
2
f
2
)
C
,1
(t
d
)]0l
2 isin
V
1,2
t
1
e
i(
f
1
2
f
2
)
C
,2
(t
d
)1l (82)
+
O
`
i=2
F
cos
V
i21,i
t
1
C
,i
(t
d
)
2 isin
V
i,i+1
t
1
e
i(
f
1
2
f
2
)
C
,i+1
(t
d
)
G
il,
where
f
1
and
f
2
are the phases (Eq. (23)) of the analysis
pulses relative to the pulses used to create the state
C
(0). We now measure the probability P
of finding the
atom in the ↓l state and obtain
P
(D
f
)=
1
/2
F
C
,0
(t
d
)
2
+ C
,1
(t
d
)
2
2 2Re[iC
*
,0
(t
d
)C
,1
(t
d
)e
i
Df
]+
O
`
i=2
C
,i
(t
d
)
2
G
, (83)
where D
f
f
1
2
f
2
. In this expression, we have as-
sumed all cross terms of the form C
*
,i
C
,i+1
for i $ 1
average to zero over many measurements; this will be
true if decoherence is caused by some process which is
uncorrelated with the creation and analysis pulses.
We fi n d kRe(C
*
,0
C
,1
)l =
1
/2kP
(p) 2 P
(0)l and
kIm(C
*
,0
C
,1
)l =
1
/2kP
(p/2) 2 P
(2p/2)l. Therefore,
from measurements of P
for four values of D
f
,we
measure the coherence (C
*
,0
C
,1
) (or
r
0,1
(t
d
), the off-di-
agonal matrix element of the motion after time t
d
). If
necessary, we can find the amplitudes C
,0
(t
d
) and
C
,1
(t
d
) by applying the blue sideband to
C
(t
d
) for a
time
t
and reducing P
(
t
) as described in Sec. 3.2.1.
4.2 Internal State Decoherence
For many years, one of the principal applications of
the stored-ion technique has been for high-resolution,
high-accuracy studies of internal state structure of
atomic ions. This capability can be applied for use in
atomic clocks (for reviews of efforts from a number of
laboratories, see Refs. [242] and [243]). High resolution
and accuracy are obtained because ions can typically be
stored for long times (many hours or days) with minimal
perturbations to their internal structure from electric
and magnetic fields.
Energy level shifts caused by electric fields (Stark
shifts) are usually small and, in many cases, magnetic
field level shifts can be controlled well enough; these
properties lead to very weak decoherence between inter-
nal state superpositions. As an example, in Fig. 7, we
show a spectrum taken of a particular hyperfine transi-
tion in
9
Be
+
ions [76]. This resonance, obtained with the
Ramsey method of separated fields [173], has a
linewidth of less than 0.001 Hz (at a frequency
v
0
/
2p . 303 MHz) and corresponds to a coherence time
between the two internal levels of the transition of more
than 10 minutes. (We could independently establish that
the noise apparent in Fig. 7 was primarily caused by the
fluctuations in the oscillator driving the transitions and
not due to decoherence of the
9
Be
+
internal levels.) More
recently, a slightly narrower resonance was reported by
Fisk et al. [244]. This resonance was observed on a
much higher frequency hyperfine transition in
171
Yb
+
(. 12.6 GHz) which resulted in a Q factor of 1.5 3 10
13
(Q transition frequency/linewidth). For levels sepa-
rated by optical energies, very long coherence times are
also possible because of the very long radiative lifetimes
of particular optical levels. So far however, observed
coherence times have been limited by the linewidth
of the probing lasers to a few tens of hertz
[235,245,246,247].
Fig. 7. Ramsey signal of the 2s
2
S
1/2
(M
I
= 2
1
/2, M
J
=+
1
/2)
2s
2
S
1/2
(2 3/2, +
1
/2) hyperfine transition in
9
Be
+
at a magnetic field of
0.8194 T). This resonance was obtained using a free precession time
of 550 s. The data are the result of one sweep (one measurement per
frequency point). The fluctuations in the data were due to the instabil-
ity of the reference oscillator used to take the spectrum. These kinds
of measurements indicated that the coherence time for superpositions
between the two hyperfine states (which could be used as qubit levels)
was longer than 10 min [76].
These relatively small rates of decoherence indicate
that quantum states considered for clock transitions are
also attractive as qubit levels for a quantum computer.
However, the long coherence times obtained in the spec-
troscopy experiments were obtained under special con-
ditions which may not always be compatible with their
use in quantum logic. Therefore, in this section, we
consider various sources of internal state decoherence
and how they might be controlled.
290
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.2.1 Radiative Decay
A fundamental limit to internal state coherence is
given by radiative decay. In general, we must consider
decay from both levels of a two-level system; for sim-
plicity, and as is often the case, we will assume the lower
state is stable.
For electric dipole radiation, the decay rate from up-
per state 2l to lower state 1l is given by
g
rad
=
v
3
0
k2
m
1l
2
/(3p
e
0
"
c
3
) where
v
0
(E
2
2 E
1
)/
"
and
m
denotes the atomic dipole operator. For magnetic dipole
radiation, we have
g
rad
=
v
3
0
k2
m
1l
2
/(3p
e
0
"
c
5
). For
hyperfine transitions, which decay by magnetic dipole
radiation, we make the approximation k2
m
1l.
m
B
,
the Bohr magneton. If we assume
v
0
/2p = 10 GHz, then
g
rad
. 10
212
s
21
, clearly small enough for the discussion
here. For optical electric dipole transitions, we make the
approximation k2
m
1l. qa
0
where 2 q is the elec-
tron charge and a
0
is the Bohr radius. If we assume
v
0
/2p =10
15
Hz (
l
= 300 nm), we find
g
rad
. 7.5 3 10
7
s
21
, which is much too fast for our purposes. However,
as is well known, a number of trapped-ion species have
first excited optical levels which are metastable. A well-
studied case is Ba
+
for which the first excited
2
D
5/2
state
radiatively decays by electric quadrupole radiation. It
has lifetime measured to be about 35 s [95,248,249,
250]. Other ions also have metastable optical levels
which are interesting for the purposes here. Some of
these are considered by Hughes et al. [63,64] and Steane
[60]. Lifetime measurements of various ions have been
compiled and reviewed by Church [251].
4.2.2 Magnetic Field Fluctuations
In the absence of purposely applied electromagnetic
fields which provide the logic operations, uncontrolled
fluctuating external magnetic fields are expected to give
the primary contribution to internal state decoherence.
Decoherence results from the fact that the energy sepa-
ration between two levels of interest depends on the
external magnetic field B. In most cases, it will be
possible to express the transition frequency between two
levels to sufficient accuracy by
v
0
+ d
v
=
v
0
+
F
v
B
G
B
0
(B 2 B
0
)
+
1
2
F
2
v
B
2
G
B
0
(B 2 B
0
)
2
, (84)
where B
0
is the average magnetic field and
v
0
is the
transition frequency for B = B
0
.
The effects of a fluctuating magnetic field depend on
the spectrum of the fluctuations. We first assume the
typical case that the fluctuations are slow enough to be
considered quasi-static during the time of a single oper-
ation
t
op
. A common source of low frequency fluctua-
tions which would typically fit into this category are
sinusoidal field fluctuations due to unbalanced currents
in ac power lines. Therefore, in the spin-
1
/2 analog, we
assume
H
internal
=
"v
0
S
z
[1 +
b
(t)], (85)
where
v
0
b
?
<< (
t
op
)
22
. For this Hamiltonian,
Schro¨dinger’s equation yields
C
(t 2 t
0
)=C
(t
0
)e
i
f
(t)
,
C
(t 2 t
0
)=C
(t
0
)e
2i
f
(t)
, (86)
f
(t)
1
2
F
v
0
t +
E
t
0
b
(t')dt'
G
.
In a sequence of logic operations,
f
(t) must be small
enough or be taken into account (see also Sec. 4.3.2).
As a second case, we assume the magnetic field varies
rapidly; specifically, we assume B varies sinusoidally:
B 2 B
0
= DBcos
v
B
t where
v
B
>> 1/
t
op
.Wefind
d
v
=
F
v
B
G
B
0
DBcos
v
B
t
+
1
4
F
2
v
B
2
G
B
0
(DB)
2
(1 + cos2
v
B
t). (87)
The term [
2
v
/B
2
]DB
2
/4 on the right side of this ex-
pression is a frequency shift which can be absorbed into
the definition of
v
0
if DB remains constant. For simplic-
ity, assume that one of the cosine terms in Eq. (87)
dominates so that d
v
sinusoidally oscillates at
v
m
=
v
B
or 2
v
B
so that
b
(t)=
b
0
cos
v
m
t in Eq. (85). The effects
of this fast modulation can be seen if we consider apply-
ing external radiation near the carrier frequency
v
0
of
the internal state transition. Schro¨dinger’s equation
leads to expressions similar to those of Eqs. (17) which
take the form (for
D
=
d
=0)
C
˙
,n
= 2 ie
i(
f
2
h
m
sin
v
m
t)
V
n,n
C
,n
,
C
˙
,n
= 2 ie
2i(
f
2
h
m
sin
v
m
t)
V
n,n
C
,n
, (88)
where
h
m
v
0
b
/
v
m
.Wehave
e
6ih
m
sin
v
m
t
= J
0
(
h
m
)
+2
F
J
2
(
h
m
)cos2
v
m
t + J
4
(
h
m
)cos4
v
m
t + ...
G
6 2i
F
J
1
(
h
m
)sin
v
m
t + ...
G
. (89)
291
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
where J
i
(X)istheith Bessel function with argument X.
For
v
m
>> 1/
t
op
.
V
n,n
, the sinusoidally varying terms
on the right average to zero (as in the rotating wave
approximation). Therefore the wavefunction evolves just
as in Eq. (25) except we must replace
V
n,n
by
V
n,n
J
0
(
h
m
). Since the J
0
factor can be absorbed into
the definition of
V
, this steady modulation should not
cause a problem. An important source of high-fre-
quency magnetic fields is from the currents in the elec-
trodes which oscillate at the trap drive frequency
V
T
.
The gate demonstrated in Ref. [17] had a strong sen-
sitivity to magnetic field fluctuations since the qubit
frequency and the auxiliary transition frequency had a
dependence
v
/B .
m
B
/
"
. 10
10
Hz/T. If this scheme
is used in future experiments, sufficient magnetic
shielding must be provided. Alternatively, the magnetic
field dependence can be minimized by operating at a
magnetic field where the transition frequency becomes
independent of magnetic field to first order (making the
v
/B terms in Eq. (84) and (87) vanish). For example,
most atomic clocks are based on hyperfine transitions in
ground electronic
2
S
1/2
states, where the nuclear spin I is
half integral (
1
/2, 3/2, 5/2, ...). In these cases, where the
total angular momentum (F = I + S) is integer, the
(F,m
F
=0)(F',m
F
= 0) transition is first-order field
independent for B. 0. These particular transitions
may not be useful in the applications discussed here
since other transition frequencies (for example,
(F,m
F
=0)(F',m
F
= 6 1) transitions) will be very
close to that of the (F,m
F
=0)(F',m
F
= 0) transition
frequency as B→0. This may cause unwanted cou-
plings to these other levels. At optical frequencies, sim-
ilar first-order field independent transitions in ions have
been used in frequency standards [235]. However, other
hyperfine and optical transitions become first-order
field independent at nonzero magnetic field, and their
frequencies are separated from competing transitions
(see, for example, Refs. [76] and [189]).
For quantum logic, storing qubit information in two
states whose energy separation is first-order field inde-
pendent therefore appears to be attractive. However, the
conditional dynamics which is central to quantum logic
may necessitate a transition to an auxiliary state. In the
notation of Sec. 3.3, a transition between the ↑l and
auxl levels will have a transition frequency which will,
in general, not be first-order field independent if the
↓l →l transition frequency is field-independent.
Therefore, if the external field fluctuates, the fidelity of
this operation will be compromised. As an example,
suppose quantum logic is performed on a collection of
9
Be
+
ions using the 2s
2
S
1/2
m
I
= 2 3/2,m
J
=1/
2l ↓l”and2 1/2,1/2l ↑l levels as qubits
(strong-field state representation). With these designa-
tions, level ↓l has a higher energy than level ↑l. These
are the same levels as those used for the clock reported
in Refs. [75] and [76]; the transition frequency
v
0
/
2p . 303 MHz is first-order field independent at a field
of B
0
. 0.8194 T. At this magnetic field, the transition
frequency has a second-order dependence of d
v
0
/
2p . 2 5.2(dB/B
0
)
2
MHz. Therefore, a field fluctua-
tion of dB =10
24
T from B
0
leads to a frequency offset
of the transition of only about 80 mHz. For the first CN
gate described in Sec. 3.3, we could use 2p transitions
between the ↑l and auxl≡1/2,1/2l level. How-
ever, this transition has a first-order dependence on
magnetic field at 0.8194 T given by (
v
/2p)/
B . 2 22 MHz/T. Therefore if the field shifts by
dB =10
24
T, this will cause a shift of this transition
frequency of
D
/2p = 2 2200 Hz. Similar consider-
ations must be applied to quantum logic using other ions
such as
201
Hg
+
ions where the splittings between hyper-
fine transitions are much higher [252]. Since it will be
difficult to find two field-independent transitions at the
same field, it may be advantageous to use logic gates
which require only one internal state transition as in the
second and third gates described in Sec. 3.3.
4.2.3 Electric Field Fluctuations
In Sec. 4.4, we treat the effects of electric fields from
the applied laser beams; in this section we treat the
effects of other electric fields. Electric field shifts are
likely to be less important as a source of internal state
decoherence than shifts due to external magnetic fields.
First, electric-field shifts are second-order in the field,
except for exceptional cases involving nearly-degenerate
states of opposite parity. Second, the fields are relatively
small at the site of the ion. In analogy to Eq. (84), we
have
v
0
+ d
v
E
=
v
0
+
1
2
F
2
v
E
2
G
E=0
E
2
, (90)
where E is the magnitude of the electric field. Unlike
Eq. (84), the expansion is made around E = 0, rather
than a nonzero value E
0
, since, as noted previously, the
mean value of E is zero for an ion in a trap. If it were not
zero, then the ion would have a mean acceleration and
would not be spatially localized. Of course, the ion must
experience an average electric field to counteract the
force of gravity, but this field is extremely small. For an
ideal rf trap, such as one in which the rf potential is
described by Eq. (1), and in which there are no stray
static electric fields, the ions are attracted to the line
along which the rf electric field is zero. In this case, the
mean-squared electric fields due to confinement experi-
enced by cooled ions are very small. If stray fields are
present, the mean-squared electric field seen by the ions
292
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
will be nonzero and can be significantly larger [228].
The stray fields can be reduced by applying potentials to
compensation electrodes; to some extent this is neces-
sary anyway for the efficient laser cooling. It should be
possible to reduce the rms electric fields from the elec-
trodes and neighboring ions to less than 1 V/cm. We
note that rms electric fields of about 8 V/cm are present
at room temperature, due to the blackbody (thermal)
electromagnetic field. Since the spectrum of this field is
dominated by infrared frequencies and because it is
steady, this should lead only to a small ac Stark shift
[253].
We briefly consider two cases that have been dis-
cussed for ion trap quantum computation: narrow opti-
cal transitions between the ground and a metastable state
and transitions between ground-state hyperfine sublev-
els. For an optical transition between two energy levels,
the energy shift, derived from perturbation theory, is
roughly (qa
0
)
2
E
2
/DE, where qa
0
is the product of the
proton charge and the Bohr radius, that is, a typical
electric-dipole matrix element, DE is a typical energy
difference between one of the two levels and another
which is connected by the electric-dipole operator, and
E is the electric field. For the case of the Ba
+
6s
2
S
1/2
-to-
5d
2
D
5/2
transition which has a wavelength of 1.76 mm,
this expression predicts a shift d
v
E
/2p of approximately
3 mHz for E = 1 V/cm, if DE is taken to be the 6s
2
S
1/2
-to-6p
2
P
1/2
energy difference. This is in rough
agreement with the experiment of Yu et al. [254]. Simi-
lar values of d
v
E
are estimated for the 5d
10
6s
2
S
1/2
5d
9
6s
22
D
5/2
transition in
199
Hg
+
[228]. Transition fre-
quencies between hyperfine levels, such as those used in
the NIST single
9
Be
+
ion experiments, are even less
sensitive to electric fields, since, to a good approxima-
tion, an electric field shifts all sublevels of the hyperfine
multiplet by the same amount. A differential shift be-
tween the sublevels arises in third-order perturbation
theory [253]. For
9
Be
+
, for example, the fractional shift
between hyperfine sublevels is approximately
2 4 3 10
218
for E = 1 V/cm [255].
So far, we have considered only electric dipole shifts,
that is, the shifts that are due to a uniform electric field.
An electric quadrupole shift, which is proportional to
the product of the electric quadrupole moment of the
state and the applied electric field gradient, may also be
present. The shift d
v
Q
is given approximately by
d
v
Q
ø
qQ
a
"
E
i
r
j
, (91)
where Q
a
is the atomic quadrupole moment and E
i
/r
j
is a typical component of the electric field gradient
tensor, such as E
x
/z. The precise value depends on the
whole electric field gradient tensor and the details of the
atomic state. There is no quadrupole shift for S-states or
for other states with electronic angular momentum
J < 1, like the
3
P
0
states in the Group IIIA ions, such as
In
+
[256]. The trap rf electric fields have gradients, but,
since they oscillate sinusoidally, they average to zero.
Static electric field gradients due, for example, to patch
effects, or the static potential to confine ions along the
axis of a linear trap [Eq. (2)] would be more of a prob-
lem. For a typical value of Q
a
=10a
2
0
, and an electric
field gradient of 10 V/cm
2
, d
v
Q
/2p 0.7 Hz. Fluctua-
tions in this (small) shift will probably be small com-
pared to those of magnetic field shifts. Electric field
gradients due to neighboring trapped ions will also be
present [257]. Consider an ion at position x = y = z =0.
The electric field gradient component E
z
/z due to ions
having charge q, located at x = y =0,z = 6 d is 2 q/
(p
e
0
d
3
). For d =10mm, this is 575 V/cm
2
. However, this
should be relatively constant, since the spacing between
ions is determined by the trap fields, which must be kept
nearly constant if motional decoherence is to be avoided.
Thus, they may lead to energy shifts, but not to decoher-
ence.
Garg [58,59] has considered the decoherence of inter-
nal states of ions due to fluctuating electric fields in-
duced by the reservoir of vibrational modes of a string
of trapped ions. Such decoherence is present even if the
modes are in the ground state, because of zero-point
motion. This form of intrinsic decoherence was found to
be negligible, at least for the specific case in which the
6s
2
S
1/2
and the 5d
2
D
5/2
states of Ba
+
are used to store the
qubits. For a string of 1000 ions, the decoherence time
was found to be around 10
4
2 10
8
times longer than the
natural lifetime of the upper state, which is 35 s.
4.3 Logic Operation Fidelity and Rotation Angle
Errors
The third and perhaps most important class of deco-
herence involves imperfections in the logic operations.
Ideally, a quantum computation transforms an initial-
ized state vector
C
(0) to U
ideal
C
(0), where
U
ideal
= exp[2 iH
ideal
t/
"
] (with H
ideal
time independent)
is the unitary evolution operator without errors. In prin-
ciple, both U
ideal
and the corresponding interaction
Hamiltonian H
ideal
operate on the Hilbert space formed
by the ↓l and ↑l states of the L qubits. In practice, for
the case of trapped ions, the physical implementation of
logic gates requires that we must include at least the 0l
and 1l states of the chosen motional mode and perhaps
an internal auxiliary state (Sec. 3.3). If decoherence
mechanisms cause other states to be populated, the
Hilbert space must be expanded. Although more
streamlined algorithms may be available, the operator
U
ideal
can always be broken up into fundamental single
and two-bit quantum gates [140,155,156,157]; for sim-
plicity, this will be assumed in what follows.
293
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
In the previous sections, we have concentrated on the
decoherence of the state vector
C
due to environmental
coupling, in the absence of purposely applied additional
fields. Here, we treat decoherence due to imperfections
in the evolution operator U
ideal
which is implemented by
these additional applied fields. We break this decoher-
ence into two categories: (i) errors or noise in the (clas-
sical) gate parameters (that is, the rotation angles
f
and
V
n',n
t in Eq. (23)) which result in undesired evolution
and, (ii) coupling to the environment induced by the
gates (for example, spontaneous emission induced by the
applied fields). In some cases, we can model noisy logic
gates by considering an ensemble of gates whose
parameters are statistically distributed. This allows a
simple characterization of several candidate noise
sources.
The final step in an ideal quantum computation is a
probabilistic measurement of all the quantum bits,
which yields a given L-bit number k, or equivalently, the
state k, with probability P
k
= kkU
ideal
C
(0)l
2
(k = 0,1,2,...,2
L
2 1). A particular quantum computation
is thus completely characterized by the vector of proba-
bility amplitudes A
k
= kkU
ideal
C
(0)l. The usefulness of
a quantum computer algorithm relies on the coherent
interference of the probability amplitudes, resulting in
only a small fraction of the 2
L
possible numbers with
appreciable measurement probability. If the actual evo-
lution operator U contains imperfections, the probabil-
ity amplitude vector is instead B
k
= kkU
C
(0)l Þ A
k
.
We characterize the fidelity of a computation by the
expression
F =
K
kU
ideal
C
(0)U
C
(0)l
2
L
=
7
O
2
L
21
k=0
A
*
k
B
k
2
8
, (92)
where the average is performed over any random vari-
ables affecting the operator U. Roughly speaking, F is
the probability that errors in the operator U do not affect
the result of a computation, with F = 1 representing
error-free computation. This characterization of the fi-
delity follows the approach of quantum trajectories
[258,259]. A quantum computation can be interpreted as
the evolution of a complicated path through Hilbert
space. If the operations comprising the computation are
imperfect, the path of computation bifurcates into a
superpositions of the correct path and incorrect path
[260]. In this model, the fidelity F is just the probability
that the correct path has been followed. Schemes for
complete characterization of a quantum process in terms
of a input-output matrix have been proposed in Refs.
[261] and [262]. This approach would be able to charac-
terize the fidelity of any of the operations discussed
here. It can also be used to test for decoherence of the
internal and motional states, by use of time delays be-
tween state preparation and application of the diagnos-
tic.
An adequate characterization of the effects of noise in
a quantum computation requires knowing the type of
noise processes present, and application of this noise to
a specific computational algorithm on a specific input
state. This is most straightforwardly accomplished using
numerical simulations [1,263–266]. Here, we do not
treat any specific algorithms, but will attempt to make
some general observations on particular sources of logic
operation errors and how these errors accumulate after
many operations.
4.3.1 Accumulated Errors
Each successive operation in a quantum computation
based on the Cirac/Zoller trapped-ion scheme can be
regarded as a rotation between two states Al
j
nl and
A'l
j
n'l in a multidimensional Hilbert space where
Al
j
and A'l
j
are from ↓l
j
, ↑l
j
,orauxl
j
and nl and
n'l are from 0l or 1l of the selected motional mode.
Before the kth operation, assumed to be applied to the
jth ion, we can, to a good approximation, write the
wavefunction as
C
k21
=
a
k21
c
k21,
a
l
j
c
¯
k21,
a
l +
b
k21
c
k21,
b
l
j
c
¯
k21,
b
l
+
g
k21
c
k21,
g
l
j
c
¯
k21,
g
l, (93)
where
c
k21,
a
l
j
and
c
k21,
b
l
j
are the basis states for the
rotation (for example, the ↑l
j
1l and auxl
j
0l states
required to carry out step 1b of Sec. 3.3). The state
c
k21,
g
l
j
is the part of the wavefunction for the jth ion
which does not include states involved in the rotation
(for example, the ↓l
j
state). The states
c
¯
k21,
a
l,
c
¯
k21,
b
l,and
c
¯
k21,
g
l are states which include other ions
and, because of gate imperfections, may involve states
outside the Hilbert space spanned by the ↓l, ↑l,
and auxl internal states, and 0l and 1l motional
states. We choose k
c
¯
k21,
a
c
¯
k21,
a
l = k
c
¯
k21,
b
c
¯
k21,
b
l =
k
c
¯
k21,
g
c
¯
k21,
g
l = 1 so that
a
k21
2
+
b
k21
2
+
g
k21
2
=
1. From Eq. (23), the “actual” kth operation can be
written as
R
ka
(
u
k
+
z
k
,
f
k
)=
cos(
u
k
+
z
k
) 2ie
i
f
k
sin(
u
k
+
z
k
)
a
k21
34
,
c
=
34
,
2ie
2i
f
k
sin(
u
k
+
z
k
) cos(
u
k
+
z
k
)
b
k21
(94)
where, to simplify notation, we have chosen
u
k
=(
V
n',n
t)
k
and
z
k
,
f
k
<< p are the errors in the ampli-
tude and phase of the rotation. We can write
294
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
R
ka
= R
k
+ DR
k
, (95)
where R
k
= R
k0
cos
z
k
[R
k0
R
k
(
u
k
, 0)] is the ideal kth op-
erator and
DR
k
has terms which are first order (and real) in
z
k
and
f
k
. Nevertheless, we might expect that the overall fi-
delity has errors which depend only quadratically on
z
k
and
f
k
since a general computation may approximate a
sequence of nearly random rotations on a multidimen-
sional Hilbert space. To see how this might come about
we express a general computation consisting of M oper-
ations as
O = R
Ma
R
(M21)a
...R
ka
...R
2a
R
1a
. O
1
+ O
2
+ O
3
, (97)
where
O
1
R
M
R
M21
...R
2
R
1
,
O
2
[R
M
...R
2
DR
1
]+[R
M
...DR
2
R
1
]...
+ ...[R
M
DR
M21
...R
1
]+[DR
M
R
M21
...R
1
], (98)
O
3
R
M
...R
3
DR
2
DR
1
+ R
M
...DR
3
R
2
DR
1
...
+ ...DR
M
R
M21
DR
M22
...R
1
+ DR
M
DR
M21
R
M22
...R
1
,
and where we have neglected terms which are higher
than second order in
z
k
and
f
k
. In this notation, the ideal
computation is given by the operator
O
ideal
= R
M0
...R
20
R
10
. (99)
For the fidelity, we need to calculate
k
C
(0)O
ideal
O
C
(0)l = k
C
(0)O
ideal
(O
1
+ O
2
+
O
3
)
C
(0)l where we will neglect terms of third order
and higher in {DR
i
}. For the first term, we find
k
C
(0)O
ideal
O
1
C
(0)l.1 2
O
M
i=1
z
2
i
2
. (100)
For the second term, we have a sum of M terms; the kth
of which is given by
k
C
(0)O
ideal
R
M
R
M21
...DR
k
...R
2
R
1
C
(0)l
.k
C
(0)R
10
R
20
...R
k0
DR
k
R
k21
...R
1
C
(0)l (101)
= k
C
k21,0
R
k0
DR
k
C
k21,0
l
where
C
k21,0
corresponds to the wavefunction repre-
sented in Eq. (93) for the ideal case. We find
k
c
k21,0
R
k0
DR
k
c
k21,0
l.2
f
2
k
2
sin
2
u
k
[
a
k21
2
+
b
k21
2
]
2 i[(
a
k21
2
2
b
k21
2
)(
z
k
cos
u
k
+ sin
u
k
)
f
k
sin
u
k
(102)
+2Re[
a
*
k21
b
k21
](
f
2
k
2
sin
u
k
cos
u
k
2
z
k
)
2 2Im[
a
*
k21
b
k21
](sin
u
k
+
z
k
cos
u
k
)
f
k
cos
u
k
].
The third term k
C
(0)O
ideal
O
3
C
(0)l is straightfor-
ward to evaluate and is clearly second order in
z
k
and
f
k
. Even though the imaginary term in Eq. (102) has
terms linear in
z
k
and
f
k
, when the fidelity F is calcu-
lated, these terms add only in quadrature so the kth
operation gives only a contribution of second order in
z
k
and
f
k
. Therefore, in general, we have
F . 1 2
O
M
k=0
F
C
k1
z
2
k
+ C
k2
z
k
f
k
+ C
k3
f
2
k
G
, (103)
where the coefficients C
k1
, C
k2
, C
k3
<
,
1. For general
computations involving many operations M in a large
Hilbert space, it might be expected that the errors in
rotations are distributed in an approximately random
fashion. In this case, we would expect the fidelity to be
approximately given by F . 1 2 M(F
zz
z
2
+ F
ff
f
2
+
F
zf
zf
) where
z
and
f
are characteristic of the errors for
all operations and F
zz
, F
ff
, F
zf
<
,
1. In the case ampli-
tude errors (
z
) dominate, then the maximum number of
operations before the fidelity drops appreciably below 1
is M
max
. 1/
z
2
. For example, laser intensity fluctuations
on the order of 1 part in 10
4
would give M
max
. 10
8
.At
the other extreme, the errors might add coherently in
some cases. These operations could conspire to give
contributions to the fidelity that goes as (
S
M
k
z
k
)
2
.We
can illustrate this with a simple example. Assume that
an intended rotation
u
k
has an error
z
k
<< p. The state of
the system before application of this rotation can be
expressed by Eq. (93). From Eq. (92) we find F =
cos
2
(
z
k
/2)(
a
k21
2
+
b
k21
2
)+4[Re{
a
*
k21
b
k21
exp(i
f
k
)}]
2
sin
2
(
z
k
/2). The existence of the second term only serves
sin
u
k
sin
z
k
i
F
(e
i
f
k
2 1)sin
u
k
cos
z
k
+e
i
f
k
cos
u
k
sin
z
k
G
DR
k
2
34
. (96)
i
F
(e
2i
f
k
2 1)sin
u
k
cos
z
k
+e
2i
f
k
cos
u
k
sin
z
k
G
sin
u
k
sin
z
k
295
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
to improve the fidelity, so at worst (when
a
j
=1 or
b
j
= 1), the fidelity of this single operation is given by
cos
2
(
z
k
/2) . 1 2 (
z
k
)
2
/8. Now suppose that a second ro-
tation
u
m
about the same axis has error
z
m
. A similar
analysis shows that when these two operations are ap-
plied in succession, a worst case for the fidelity is given
by F = cos(
z
k
/2 +
z
m
/2) . 1 2 (
z
k
+
z
m
)
2
/8. Similarly,
for rotations on M bits, the worst case fidelity is given
by F = cos(
S
M
i
z
i
/2) so that the errors could accumulate
linearly. Most likely, this case will correspond to rather
uninteresting computations and one can generally expect
that the errors will accumulate in a way which is be-
tween random walk and linear.
The effects of operation errors (
z
i
,
f
i
) have been
observed in simulations of quantum computations to
factorize small numbers [263,266]. If the errors fluctu-
ate randomly about the correct value, the fidelity is
given by a simple expression which is essentially the
same as Eq. (103) for F . 1 [266]. However, for con-
stant errors the fidelity indicated by the simulations is
somewhat worse indicating that the computational al-
gorithms cannot be assumed to be rotations about axes
chosen pseudo randomly. Therefore, the simulations in-
dicate that requirements on systematic offsets are more
stringent than the requirements on random fluctuations
from operation to operation.
4.3.2 Pulse Area and Phase Fluctuations
As in the last section, we will consider two sources of
noise: (1) those which result in offsets from, and fluctu-
ations in, the ideal Rabi frequencies [for example, off-
sets and fluctuations of
V
n',n
t in Eq. (21)] and (2) those
caused by phase fluctuations between applications of
separate operations to a particular ion [fluctuations from
some constant value of
f
in Eq. (21)].
We first make some remarks regarding the coherent
evolution between two states of the system which com-
prise the basis states for a single operation. First, there
are a number of methods of coherently transferring pop-
ulation between a two level system with radiation, in-
cluding constant amplitude Rabi pulses as represented
by Eqs. (21) and (23), adiabatic rapid passage [267], and
adiabatic transfer via dark states [268–270]. The advan-
tage of the last two techniques is that for p pulses, their
efficiency does not depend critically upon the pulse area
(Eq. (104) below) of the applied field. However, for
rotations other than p, this advantage is lost. Moreover,
for these techniques to work with high efficiency, par-
ticularly simple atomic level structures are required and
coupling to any off-resonant levels can potentially con-
taminate the transfer. For example, adiabatic transfer
can be extremely efficient between ground state hyper-
fine levels if the transfer occurs only through a single
excited (electronic) state. However, in practice, other
excited states are driven by the coupling (laser fields)
thereby affecting the fidelity of the transfer [271]. Fur-
thermore, the high powers required for adiabatic rapid
passage and adiabatic transfer aggravate decoherence
from spontaneous Raman scattering (Sec. 4.4.6.4).
Therefore, we will only consider the case of Rabi
pulses, where population is transferred by applying a
resonant field with a well-controlled amplitude envelope
and duration.
When a resonant radiation field is applied to a two-
level system, the resultant unitary transformation is
given by Eq. (23). We can further generalize this expres-
sion replacing
V
n',n
t by
Q
n',n
(t)/2 where
Q
n',n
(t)=2
E
t
t'=0
V
n',n
(t')dt' (“pulse area”). (104)
The factor of two in this equation is introduced so that
the condition e
V
n',n
(t')dt' = p/2 corresponds to a p pulse
of the effective Bloch vector. Here, as opposed to the
situation assumed in Sec. 2.3.1 (where the coupling
radiation was assumed to be turned on instantaneously,
left at constant amplitude, and then turned off instanta-
neously), we now allow the coupling radiation to be
turned on and off gradually. This more closely approxi-
mates a real experimental situation and can be advanta-
geous since the spectrum of
V
n',n
(t) can be tailored to
suppress off-resonant transitions (Sec. 4.4.6). Here, we
call
Q
n',n
the pulse area of the applied radiation and
2
V
n',n
(t) the Rabi frequency envelope. If the applied
pulse area is noisy or set incorrectly, then the output
state will fluctuate from it’s ideal value as discussed in
the previous section.
We must also consider detunings between the applied
frequencies and resonant frequencies of the ions [in Eq.
(21)] and phase fluctuations between successive opera-
tions on the same ion [fluctuations in
f
in Eq. (21) from
operation to operation]. This could occur from fre-
quency or phase fluctuations in the applied fields or
fluctuations in the ion internal or motional frequencies.
For frequency fluctuations, we consider that for a given
operation on ion j,
D
j
(t) is slowly varying over the time
of a single operation (Sec. 4.2.2). In the limit that detun-
ings are small, the dominant effect can be characterized
by assuming that
D
= 0 for each operation, but that
different phases
f
k
=
E
t
k
t
0
D
j
(t')dt', (105)
are present for successive applications of the same oper-
ation. In this expression, t
0
is the time that the first
operation is applied at ion j and t
k
is the time of the kth
operation (applied to the jth ion).
296
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
4.4 Sources of Induced Decoherence
Below, we consider various sources of fluctuations
and decoherence caused by the logic operations and how
these might be evaluated and controlled in experiments.
First, we are interested in controlling the accuracy and
stability of the Rabi pulse are
Q
n',n
given by Eq. (104)
and the accumulated phase as expressed in Eq. (105).
Since gates will most likely be implemented with laser
transitions we will examine laser intensity and timing
fluctuations.
4.4.1 Applied Field Amplitude and Timing
Fluctuations
Fluctuations in the laser intensity at the site of a given
ion can arise from both fluctuations in the relative posi-
tion of the beam with respect to the ion and fluctuations
in laser power. Laser/ion position stability is particularly
important since the Cirac/Zoller scheme of quantum
logic assumes that ions in an array be selectively ad-
dressed, thereby requiring a high degree of control of
the laser beam spatial profile (Sec. 4.4.4). Of course, the
simplest method for minimizing position fluctuations is
to employ mechanically stiff mounts for the optics and
ion trap electrodes, and have the laser source as close as
possible to the ions. A quadrant detector indexed to the
trap electrodes and placed near the ion may also be used
to actively stabilize the beam position by feeding back to
a galvanometer or acousto-optic modulator. If optical
fibers are used to deliver laser beams to the ions, posi-
tion fluctuations between the fiber and the ions could be
made small; however, we must also consider position
fluctuations between the laser source and the input to
the fiber. If the position tolerances can be adequately
controlled, the dominant source of intensity noise at the
ion would likely be given by fluctuations in optical
power and laser mode. Here, we estimate limits on laser
amplitude noise.
If we assume the laser fields responsible for quantum
logic operations are coherent states, the fundamental
noise floor is photon shot noise. For a laser beam of
average power P
0
, the fractional level of shot noise is
dP
P
0
= Î
"v
P
0
t
op
, (106)
where
v
is the (optical) photon frequency,
t
op
is the time
the radiation is applied, and, for simplicity, we assume
square pulse envelopes as in Eq. (23). Of course, lasers
seldom produce amplitude noise at the photon shot-
noise limit. Almost all laser sources have significant
amplitude noise in the 10 Hz to 10 kHz range due to
acoustic vibrations which, for example, affect the laser
cavity resonators. Much of this noise can be removed by
active power stabilization, where a beamsplitter directs
a portion of the laser power to a photodetector, and an
error signal is derived and fed back to an upstream
modulator or, in the case of a diode laser, directly to the
amplitude of the laser source [272,273]. The limiting
noise of this stabilization scheme is degraded slightly by
the imperfect quantum efficiency of the photodetector
as well as the beamsplitter. If the beamsplitter directs a
fraction
e
of the input optical power to the stabilizer
(which then gives an optical power P
u
. (1 2
e
)P
0
di-
rected to the ion), and the quantum efficiency of the
photo detector is
h
det
, the limit of fractional power noise
in the logic pulse is (assuming no added electronic noise
in the feedback loop)
dP
u
P
u
$ Î
"v
P
u
t
op
h
det
e
(1 2
e
)
. (107)
For a laser wavelength of 313 nm, and assuming
e
= 0.5
and
h
det
= 0.5, we have dP
u
/P
u
$ 2.3 3 10
29
(P
u
t
op
)
21/2
.
For1Wofusable laser power and
t
op
=1 ms, this
corresponds to a fractional power fluctuation of
$ 2.3 3 10
26
.
This estimate applies only to the laser power fluctua-
tions at the beamsplitter and assumes no additional noise
is introduced between the beamsplitter and the photo
detector or the beam splitter and the ions. Typically, the
usable part of the laser beam must be directed further
through optics, the air, and a window to the vacuum
envelope enclosing the ion trap. Fluctuating etalon ef-
fects in the optics and air currents may therefore seri-
ously increase the power fluctuations beyond Eq. (107).
The effects of (Gaussian) noise on laser intensity have
been treated by Schneider and Milburn [274]. These
effects show up in a well-characterized way for transi-
tions involving Fock states.
Fluctuations in timing errors may also be important.
Clearly, fractional fluctuations in the duration of laser
pulses will correspond directly to the same fractional
fluctuations in the desired value of
Q
n',n
. If we require
fractional fluctuations of 10
26
on
Q
n',n
, then we require
timing precision of 1 ps on a 1 ms pulse. Similar consid-
erations apply to the stability of pulse envelope shapes.
For both amplitude and timing fluctuations, it may be
possible to sample a portion of the beam used for logic
and apply it to a “check bit” ion. The response of this ion
could then be used to monitor and control the amplitude
and timing of the pulses.
4.4.2 Characterization of Amplitude and Timing
Fluctuations
In practice, it may be easiest to characterize the am-
plitude and timing fluctuations with a power detector
297
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
placed as close as possible to the position of the ions.
However, it would be useful to characterize these effects
using the ions themselves. One possibility is to observe
the sinusoidal Rabi oscillations predicted by Eq. (21).
For example, for a single ion initially prepared in the ↓l
state, we can record the probability P
(
t
) of detecting
the ion in the ↓l state after the laser is applied to the ion
for time
t
. From Eq. (23), if we first prepare the ion in
the ↓,nl state and if the power is constant,
P
(
t
)=C
,n
2
= cos
2
V
n',n
t=
1
/2[1 + cos2
V
n',n
t]. In
Fig. 2, we show this type of curve for n = 0 and n' =1.
This plot includes a decay due to decoherence. Here, we
consider decoherence caused by specific types of noise.
We will cast the results in terms of fluctuations in
V
n',n
,
however, they may easily be converted to corresponding
time fluctuations since we are really interested in fluctu-
ations in the net rotation angle given by Eq. (104). For
simplicity of notation, in this section, we write
V
n',n
=
V
.
First assume that
V
fluctuates slowly so that it is
constant over the time
t
for an individual measurement
of P
(
t
) but fluctuates over the time taken to make an
average measurement of kP
(
t
)l. We characterize power
fluctuations by a spectrum D(
V
2
V
0
)of
V
values
around the desired value
V
0
.Wehave
kP
(
t
)l =
1
/2
E
`
2`
[1 + cos2
Vt
]D(
V
2
V
0
)d(
V
2
V
0
).
(108)
We illustrate with two possibilities for D, finding
kP
(
t
)l =
1
2
F
1 + cos2
Vt
e
22(D
V
rms
t
)
2
G
for
D(
V
2
V
0
)=
1
Ï2pD
V
rms
e
2
1
2
S
V
2
V
0
D
V
rms
D
2
,
(109)
kP
(
t
)l =
1
2
F
1 + cos2
Vt
1
1+2(D
V
rms
t
)
2
G
for
D(
V
2
V
0
)=
1
D
V
rms
e
2
Ï2
V
2
V
0
D
V
rms
where (D
V
rms
)
2
k(
V
2
V
0
)
2
l. In both cases the contrast
of the signal drops about a factor of two for
t
. 1/
D
V
rms
. Although the second choice of D(
V
2
V
0
)
seems less physical, it indicates the exact spectrum of
the noise may not be important for a qualitative test. If
we need to detect fluctuations of order D
V
rms
/
V
, then
we must measure P
after approximately
Vt
/2p .
V
/(2pD
V
rms
) Rabi cycles. For example, if we require
sensitivity D
V
rms
/
V
. 10
24
, we need to make measure-
ments of P
(
t
) after about 1500 Rabi cycles.
If, instead, the intensity fluctuates slowly compared
to
t
op
, but fast compared to the time
t
, kP
(
t
)l isaless
sensitive test of the fluctuations since their effects tend
to average out. In order to show the effects of high-fre-
quency fluctuations of
V
for a simple case, we consider
a sinusoidal time variation:
V
(t)=
V
0
+ D
V
sin(
v
amp
t +
w
), (110)
v
amp
is the modulation frequency of the Rabi amplitude
V
(t). This analysis should, at least, yield the dependence
on the amplitude and frequency of the fluctuations. The
measured quantity P
(
t
)isgivenby
P
(
t
)=
1
2
[1 + cos
u
(
t
)], (111)
where the rotation angle
u
(
t
) is given by Eq. (104)
which, here, takes the form
u
(
t
)
2
=
E
t
0
V
(t)dt =
V
0
t
+ D
V
E
t
0
sin(
v
amp
t +
w
)dt
(112)
=
V
0
t
+
D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
].
This yields
P
(
t
)=
1
2
F
1 + cos
S
2
V
0
t
+
2D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
DG
=
1
2
F
1 + cos(2
V
0
t
)cos
S
2D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
D
(113)
2 sin(2
V
0
t
)sin
S
2D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
DG
.
298
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
Since P
(
t
) is obtained by repeating the measurements,
and since the phase
w
will be random from one measure-
ment to the next, we average over
w
:
If we assume that D
V
/
v
amp
<< 1, which is valid for a
small, high-frequency, modulation, then the integrands
can be expanded to second order in (D
V
/
v
amp
):
Aside from the fast oscillations due to the cos
v
amp
t
term, which might be difficult to observe, the effect of
the high-frequency fluctuations of
V
is to reduce the
visibility of the signal by 2(D
V
/
v
amp
)
2
. This result can
be averaged over a distribution D
V
(
v
amp
).
4.4.3 Applied Field Frequency and Phase
Fluctuations
A simple source of fluctuations is due to frequency or
phase fluctuations in the radiation that is used to carry
out the basic operations. Schneider and Milburn [274]
have considered a specific model of phase fluctuations
for ion experiments and calculate the corresponding
decoherence for operations of the type used in quantum
logic. If qubit energy levels are separated by optical
energies, the lasers that drive qubit transactions must
have the required frequency and phase stability. Given
the performance of current stable lasers [235,245,246,
247], this may be a problem for long computations. As
discussed in Sec. 2.3.3, optical stimulated-Raman trans-
actions provide the required strong field gradients
whereas the overall frequency sensitivity in the transi-
tions is dominated by the difference frequency of the
lasers, not the frequency of each laser. Since the two
Raman beams can be derived from one laser beam with
the use of frequency modulators, the frequency fluctua-
tions are dominated by those of the oscillator which
drives the modulator. The phase stability of these
sources can be high and does not appear to pose prob-
lems. For example, the data shown in Fig. 7 can be
kP
(
t
)l
w
=
1
2
+
cos(2
V
0
t
)
4p
E
2p
0
cos
S
2D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
D
d
w
(114)
2
sin(2
V
0
t
)
4p
E
2p
0
sin
S
2D
V
v
amp
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
D
d
w
.
kP
(
t
)l
w
1
2
+
cos(2
V
0
t
)
4p
E
2p
0
F
1 2
1
2
S
2D
V
v
amp
D
2
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
2
G
d
w
2
sin(2
V
0
t
)
4p
E
2p
0
S
2D
V
v
amp
D
[cos
w
(1 2 cos
v
amp
t
) + sin
w
sin
v
amp
t
]
G
d
w
(115)
=
1
2
+
cos(2
V
0
t
)
2
F
1 2 2
S
D
V
v
amp
D
2
(1 2 cos
v
amp
t
)
G
.
regarded as one test of the phase sensitivity of the
oscillator used to drive the indicated internal state tran-
sition. Stimulated-Raman transitions have the disadvan-
tage that they give rise to ac Stark frequency shifts as
indicated by the g
1,2
2
/
D
R
terms on the right side of
Eqs. (40). If these shifts are equal for both qubit levels,
the net shift is zero. If they are different, the effective
qubit frequency is shifted during the operation. This
must be measured and accounted for. The shift can be
tuned to zero by appropriately adjusting
D
R
and the
relative intensities and polarizations of the Raman
beams. A problem still exists if the Stark shifts fluctu-
ate; we examine the consequences of those fluctuations
here.
From Eqs. (40), the ac Stark frequency shift of the
qubit transition due to stimulated-Raman transition is
v
S
=
v
S2
2
v
S1
= 2 (g
2
2
2 g
1
2
)/
D
R
, (116)
so that the shifted qubit frequency is given by
v
'
0
=
v
0
+
v
s
. We first consider the effects of frequency
fluctuations of the Raman beams. If the two beams are
derived from the same beam with the use of a modula-
tor, the frequency shift of the two laser beams will be
the same. Therefore, to a good approximation, we find
d
v
'
0
. 2 (d
D
R
/
D
R
)
v
s
where d
D
R
represents the fre-
quency shift of both laser beams. Typically,
D
R
>10
GHz, and frequency fluctuations of lasers can be con-
trolled to less than 1 kHz, so this source of error should
be small. Moreover, it is essentially absent if we tune
v
s
to be zero.
299
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
More serious are fluctuations in laser intensity. We
will characterize these fluctuations by the correspond-
ing fluctuations in g
i
(~ square root of laser intensity).
From Eq. (116), we have
d
v
S
=2
v
S2
dg
2
g
2
2 2
v
S1
dg
1
g
1
. (117)
As a worst case, we will assume fluctuations in g
1
and
g
2
are random and uncorrelated and, for simplicity, we
assume k(dg
1
/g
1
)
2
l.k(dg
2
/g
2
)
2
l
j
2
, where kl denotes
an average over the spectrum of the fluctuations. There-
fore
k(d
v
S
)
2
l =4
j
2
(
v
2
S1
+
v
2
S2
). (118)
As discussed in Sec. 4.3.2, we assume the fluctuations
are slow compared to
t
op
, where
t
op
is the time of the
operation. In a particular operation, a frequency offset
d
v
S
gives rise to a phase fluctuation d
f
S
= d
v
S
t
op
.Itis
useful to characterize the effects of these phase fluctua-
tions by comparing their size relative to fluctuations in
rotation angles caused by the corresponding fluctuations
in the Rabi rates
V
n',n
. With the same assumptions re-
garding dg
1
and dg
2
,wehavek(d
V
n',n
/
V
n',n
)
2
l =2
j
2
.
Therefore, fluctuations in the overall rotation angle
Q
n',n
=2
V
n',n
t
op
[Eq. (104)] are given by k(d
Q
n',n
/
Q
n',n
)
2
l =2
j
2
. From these expressions we find
k(d
f
S
)
2
l
k(d
Q
n',n
)
2
l
=
(
v
2
S1
+
v
2
S2
)
2
V
2
n',n
. (119)
For quantum logic, a worst case appears to be for side-
band excitation n' = n 6 1. If we assume the Lamb-
Dicke limit, we find
k(d
f
S
)
2
l
k(d
Q
n',n
)
2
l
=
(g
4
1
+ g
4
2
)
2
h
2
g
2
1
g
2
2
. (120)
From this expression, it is desirable to have g
1
. g
2
to
minimize the effects of d
f
s
relative to d
Q
n',n
.Evenin
this case, rms phase fluctuations caused by Stark shifts
are worse than those caused by Rabi frequency fluctua-
tions by a factor of . 1/
h
. Most likely, however, fluctu-
ations in intensity will be dominated by fluctuations
from the primary laser from which both beams are
generated. Therefore we expect fluctuations to be corre-
lated dg
1
/g
1
= dg
2
/g
2
. When, in addition,
v
S1
.
v
S2
, the
phase fluctuations caused by Stark shifts will be less
than those caused by Rabi frequency fluctuations. If
these conditions hold true, it appears that stimulated
Raman transitions between two qubit levels separated
by fairly low (for example, hyperfine) frequencies are
superior to single-photon transitions between qubit lev-
els separated by optical frequencies. This situation
might change as laser oscillators become more stable.
In the above discussion, we have neglected the alter-
ation of the ion motional frequencies caused by the
superimposed dipole force potentials of the focussed
laser beams. However, if we assume the Raman beams
have waists w
0
which are approximately equal and that
the magnitudes of the projections of their k vectors
along the z axis are the same, it is straightforward to
show that the relative shift of (single trapped ion) secu-
lar frequencies for states ↓l and ↑l is approximately
equal to d
v
S
(z
0
/w
0
)
2
. Therefore, since z
0
<< w
0
, the fre-
quency shifting effects will be dominated by the shifts
of the internal states.
Phase fluctuations of the laser beams themselves will
also directly affect the fidelity of the operations. In the
stimulated-Raman case, where both beams are derived
from the same beam, phase fluctuations between Ra-
man beams will be very nearly cancelled. However, an-
other source of laser phase fluctuations will come from
path length fluctuations between the (laser) source and
the trapped ions. Path length fluctuations are expected
to be dominated by mechanical vibrations; these vibra-
tions are typically restricted to low frequencies (< 1
kHz). They could result from a number of causes such
as fluctuating mirror mounts or trap mounting hard-
ware. For single photon laser transitions, the overall
path length between the laser and ions is important; for
stimulated-Raman transitions, the primary problem will
be caused by path length differences between the two
Raman beams after the frequency modulator. Therefore,
for stimulated-Raman transitions using overlapping, co-
propagating beams (which can be used to drive carrier
n' = n transitions) the paths are the same and there
should not be a problem. However, for stimulated Ra-
man sideband transitions, we require k
1
and k
2
along
different paths, so that the path length problem is
analogous to the problem for single photon transitions.
For brevity, we will treat the problem of single-photon
transitions; other cases (including the stimulated-Ra-
man case) are easily generalized from this.
We assume the laser electric field at the exit of the
laser oscillates as cos(
v
L
t). At the position of the ion,
the field is given by the same expression with the time
replaced by the retarded time t 2 d
L
/c where d
L
is the
distance between the laser and ion. Therefore, the overall
phase difference between the field at the laser and ion is
v
L
d
L
/c. We will assume the fluctuations in d
L
are slow
enough that over the time of a single operation the phase
can be considered constant. If d
L
fluctuates by an
amount Dd
L
, the phase of the field fluctuates by Dd
L
v
L
/
c =2pDd
L
/
l
where
l
is the laser wavelength. In the
NIST single
9
Be
+
ion experiments, all of the operations
have typically taken less than 1 ms; therefore vibrations
300
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
have not been a problem. However, in longer computa-
tions, the requirements on Dd
L
are stringent. Some form
of active stabilization, such as the method described by
Bergquist et al. [235], will probably have to be used.
That technique can be viewed as a Doppler shift cancel-
lation scheme and was based on a Doppler shift cancel-
lation scheme used in spacecraft tracking [275]. To il-
lustrate what is, in principle, possible, we note that
optical cavities can be made to track the frequency of
lasers to precision much smaller than 1 Hz
[235,245,246,247]. For a laser frequency of 5 3 10
14
Hz (
l
= 600 nm), and a cavity composed of two mirrors
separated by a distance of 50 cm, this corresponds to
holding relative positions of the mirror to within 10
213
cm. Therefore, although they add additional complica-
tions to the experiments, such schemes for length stabi-
lization can be used to effectively null the effects of path
length variations.
4.4.4 Individual Ion Addressing and Applied Field
Position Sensitivity
The scheme of Cirac and Zoller for trapped ion quan-
tum logic requires that ions (along the axis of a linear
trap) be addressable individually with laser beams for
logic operations. This may be difficult, because the high
vibrational frequencies desired for efficient laser cool-
ing and suppression of decoherence also results in
closely spaced ions. As discussed in Sec. 2.1, the mini-
mum separation of adjacent ions in a linear trap is be-
tween the center ions and is approximately
s
min
. 2sL
20.56
, with s =(q
2
/4p
e
0
m
v
2
z
)
1/3
where L is the
number of ions. For
9
Be
+
ions with an axial COM fre-
quency of
v
z
/2p = 1 MHz, this separation is about 4 mm
for 10 ions.
The most straightforward method for individual opti-
cal addressing is to tightly focus laser beams on the
selected ion [1]. The transverse intensity distribution of
a Gaussian optical beam of power P is
I(r)=
2P
pw
2
0
exp
S
2
2r
2
w
2
0
D
, (121)
where w
0
.
l
/(p?NA) is the beam waist,
l
the radiation
wavelength, and NA = tan
u
is the numerical aperture of
the beam with cone half-angle
u
(the formula for w
0
in
the paraxial ray approximation is valid only for NA <1)
[276]. For large numerical apertures (NA 0.5), beams
can apparently be focussed to w
0
.
l
, but this is diffi-
cult to realize in the laboratory. If w
0
=5mm in a Gaus-
sian beam, at a distance 10 mm from the center of the
beam, this would imply a relative intensity of about
3 3 10
24
(relevant for stimulated-Raman transitions) or
an electric field amplitude (relevant to single photon
transitions) of 1.8 % relative to the center of the beam—
clearly a problem. If w
0
=2mm could be obtained, the
intensity (electric field) would be down by a factor of
1.3 3 10
214
(1.1 3 10
27
). However, these estimates are
too optimistic because they assume the laser beam is
normal to the axis of the linear trap; to address the axial
modes by the methods described in Secs. 2 and 3, we
need a component of the laser beam k vector along the
axis. Therefore, if the angle of the laser beam relative to
the trap axis is
f
k
, we must replace r in Eq. (121) by
s
min
sin
f
k
. In addition, imperfections in the surfaces of
the intervening vacuum port window, multiple reflec-
tions from these windows, and diffraction typically dis-
tribute laser intensity outside of the theoretical waist of
the beam. The degree to which this occurs depends on
the details of window surfaces, etc. and must be resolved
experimentally.
Alternatively, we could accomplish all of the opera-
tions described in Secs. 2 and 3 by using the transverse
gradient of the field associated with a focused beam
[125]. Therefore to accomplish sideband transitions, we
would displace the laser beam laterally, with respect to
its k vector from the position of the ion. This would give
rise to a coupling to the ion’s motion which is in the
direction of the transverse intensity gradient of the laser
beam. In this case, we can make the direction of the
beam normal to the trap axis. This method has the
disadvantage that the field gradient would be reduced
relative to the case treated above thereby leading to
smaller Lamb-Dicke parameters and correspondingly
reduced values of
V
n',n
for n' Þ n.
The transverse intensity gradients of focused laser
beams can also cause significant intensity fluctuations
at the selected ion if the relative position of the beam
with respect to the ion is not stable on the time scale of
the computation. An alternative to using tightly focused
Gaussian laser beams is to first feed the (expanded)
laser beam through a sharply defined aperture, and use
a lens to image the aperture at the position of the ions.
With this technique, the beam intensity can be dis-
tributed more smoothly around the selected ion and have
very steep intensity edges (on the order of the original
aperture sharpness) away from the ion, thus suppressing
beam vibration problems and confining the radiation to
a single ion. This technique has been used to make
relatively “hard” walls for an optical dipole trap [277].
For this technique to work well, the imaging lens must
collect a large fraction of the light transmitted through
the aperture or else diffraction effects will result in light
intensity outside the image of the aperture. To address
individual ions, we require very small aperture images,
which gives rise to a design tradeoff. If a one-to-one
relay lens is used to image a small object aperture,
effects of diffraction are enhanced. If a demagnifying
301
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
lens is used to reduce a large object aperture, then the
aperture must be placed a large distance from the lens,
requiring a relatively large lens. For two ions, imaging
a sharp edge such as a razor blade at the space between
the ions may be sufficient. We might also consider hav-
ing every other ion in a string be a “garbage” ion which
is not used in the computation, thereby increasing the
spacing between qubit ions by a factor of two (or more,
if more garbage ions are used between each qubit ion).
This has the disadvantage that total number of ions (and
spectator modes) increases, aggravating the problems
associated with large quantum registers. If sufficiently
good addressing on one ion in a string can be accom-
plished, it may be simpler to adiabatically shift the posi-
tion of the ions, rather than the laser beams, in order to
address different ions. This could be accomplished by
applying different static potentials U
0
(t)andU
'
0
(t)tothe
end segments of the rods in Fig. 1. However, changes in
U
0
(t)andU
'
0
(t) would have to be coordinated to keep the
COM axial frequency constant or else additional phase
shifts would be introduced. Stimulated-Raman transi-
tions have the advantage that Dk can be made parallel to
the axis of the trap even though each beam is at an angle
with respect to the trap axis to facilitate selection of a
particular ion. This is important, since coupling to trans-
verse modes is eliminated.
Another method of optically addressing individual
ions is to cause a destructive optical interference at the
position of a specific ion, with a net coupling at the
other ion(s). For instance, if ion j is positioned at the
node of a resonant standing wave laser field, the cou-
pling between states ↓l
j
n
k
l and ↑l
j
n
k
'l is propor-
tional to kn
k
'sin[
h
j
k
(a
k
+ a
k
)]n
k
l. In this case, the cou-
pling of the standing wave to ion j vanishes when the
laser frequency is tuned to an even order sideband such
as the carrier (n
k
' = n
k
) (also see Sec. 4.4.6.2). If, in-
stead, the ion is positioned at an antinode, the coupling
is proportional to kn
k
'cos[
h
j
k
(a
k
+ a
k
)]n
k
l; thus, the
coupling vanishes when the laser frequency is tuned to
an odd order sideband, such as the first blue or red
sideband (n
k
' = n
k
6 1). By appropriately choosing the
angles of focused laser beams relative to the trap axis or
the spacing between ions, it should be possible to posi-
tion an antinode (node) at ion j while positioning nodes
(antinodes) at the ions adjacent to ion j (for equally
spaced ions). In the case of two-photon stimulated Ra-
man transitions, we desire to place ion j at common
nodes or antinodes of two standing waves. Although this
interference technique should allow individual access to
each of two trapped ions, it appears technically difficult
to extend this technique to more than three ions.
Finally, we consider the application of external field
gradients which shift the internal energy levels of ions
depending on their position. For a magnetic field gradi-
ent to give this selectivity, we require the Zeeman split-
ting between adjacent ions to be much larger than the
Rabi frequency, or D
m
(B/z)s/
"
>>
V
, where D
m
is
the difference in k
m
?Bl/B between the two levels of
interest, and s is the ion separation along the z direction.
For D
m
=
m
B
, s =10mm, and a Rabi frequency of
V
/
2p 1 MHz, this requires B/z >> 0.1 T/cm. Field
gradients of this magnitude can be achieved; however,
they would introduce large, and not-easily-controlled
phase shifts for the other ions in a quantum register.
The laser beam itself can provide ion selectivity by
employing the transverse gradient in the optical field
intensity. For instance, if we desire to perform a
u
-pulse
on ion j without affecting neighboring ion k, the inten-
sity profile of the laser beam can be set so that the ratio
of field strengths (intensities for the case of two-photon
stimulated-Raman transitions) at ion j vs. ion k is
u
/
2pm, where m is an integer. Now if the pulse duration
is set so that ion j is rotated by
u
,ionk receives a rotation
of 2pm and hence returns back to its initial state (with
an extra phase factor of (21)
m
).
For the case of two-photon stimulated-Raman transi-
tions, the laser beam can provide ion frequency selectiv-
ity by employing the Stark shift and the transverse gra-
dient of the optical field. Here, for example, we could
assume that two counterpropagating Raman beams of
equal intensities and spatial profiles are offset so that
beam 1 is centered on ion j, and beam 2 is centered on
adjacent ion k as depicted in Fig. 8. We will assume the
coupling scheme of Fig. 3. Let
e
be the fraction of peak
intensity seen by the offset ions (that is, the intensity of
beam2ationj and beam 1 at ion k). Assume that when
both beams are positioned on either ion, g
1
= g
2
= g.
When the beams are offset, the two-photon resonant
Rabi frequency at each ion is
V
=
e
1/2
(g
2
/
D
R
), where
g
2
/
D
R
is the Rabi frequency expected if both beams were
centered on a given ion. The Stark shifts of the two ions
are in opposite directions:
d
j
=+
d
0
,
d
k
= 2
d
0
, where
d
0
=
V
(1 2
e
)/
e
1/2
.Ifwemake
d
0
>>
V
(
e
<< 1), then
by appropriately tuning the difference frequency of the
laser beams, we can selectively drive transitions on ei-
ther ion j or k. Alternatively, if, for example, we desire
to perform a
u
-pulse to ion j without affecting ion k in
an “unrepairable” way,
e
can be adjusted to a particular
value which results in ion k making the transformation
c
k
=
a
↓l
k
+
b
↑l
k
c
'
k
=e
i
z
(
a
↓)
k
+e
i
j
b
↑l
k
). For
example, if we use Rabi pulses which have a square
pulse envelope, Eq. (21) applies. For ion j, the laser
must be tuned to the shifted frequency
v
'
j0
=
v
j0
+
d
0
;
Eq. (21) applies with
D
= 0 and
V
n',n
t =
u
/2 where t is
the pulse time. For ion k,
v
'
k0
=
v
k0
2
d
0
; Eq. (21) ap-
plies with
D
=2
d
0
. To achieve the desired form of
c
'
k
,
we want
u
/2(1 +
d
2
/
V
2
)
1/2
= mp,or
e
2
2 [1+(2np/
u
)
2
]
e
+1=0, where m is an integer. For m = 1 and
302
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
Fig. 8. Scheme for differential ac Stark shifting of neighboring ions. Equal intensity counter propagating beams
1 and 2 are centered on ions j and k, respectively. A fraction
e
of the peak intensity I
0
of each beam is applied
to the other ion. This results in a differential ac Stark shift of ions j and k, allowing the possibility of individually
accessing the ions by tuning the frequency of the laser beams.
u
= p (a p-pulse on ion j), we find
e
= 0.208, and
j
= 1.74p. The phase shift
j
on ion k and the corre-
sponding phase shift on ion j must be kept track of in
subsequent operations on these ions. Generalizing this to
more than two ions becomes difficult if the laser beams
also overlap other qubit ions. This scheme places an
additional premium on laser power stability, since the
light shifts are bigger than the Rabi frequencies by 1/
e
1/2
for
e
<< 1. In addition, in both of the above schemes
employing the laser beams to differentially address
neighboring ions, one major drawback is that the posi-
tions and profiles of the laser beams must be accurately
controlled.
Many of the above differential addressing schemes
appear useful when dealing with only two ions instead
of a string of many. This leads us to seriously consider
systems where quantum logic operations are performed
on accumulators consisting of only two ions, with the
other ions located somewhere else (Sec. 5.1). A pro-
posal has also been made for transferring quantum in-
formation from one register to another by optical means
[151,278].
4.4.5 Effects of Ion Motion (Debye-Waller
Factors)
The Rabi frequency
V
n',n
describes the transition rate
between states ↓lnl and ↑ln'l. To realize the condi-
tional dynamics desired for quantum logic, we want
V
n',n
to depend on the motional quantum numbers n and n' of
a particular vibrational mode and be independent of the
state of motion of other modes. In addition, for simple
rotations on internal states, we want
V
n',n
to be indepen-
dent of the motional state for all modes. It is not possi-
ble, in practice, to rigorously satisfy both of these
requirements. For instance, the conventional controlled-
not gate employs two carrier pulses (steps (1a) and (1c)
in Sec. 3.3) which are intended to not depend on the
state of motion; this requires the Lamb-Dicke parameter
h
to be small [see Eq. (56)]. In the Raman configura-
tion,
h
is proportional to the difference in two wavevec-
tors and can be made negligible by using co-propagating
beams (Dk . 0). On the other hand, with single-photon
optical transitions, the Rabi frequencies depend on the
motion of all modes which have a component of motion
303
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
along the direction of k. As discussed in Sec. 2.3, we
can take advantage of the motional dependence of the
carrier to construct a logic gate, but in this case also, the
Rabi frequency will depend on the motion in the other
modes along the direction of k or Dk. Similarly, for
sideband operations, such as step (1b) in Sec. 3.3, it will,
in general, be impossible to have
V
n',n
, depend on only
one mode of motion. In this section, we examine the
influence of extraneous modes on the Rabi frequencies
V
n',n
.
Transitions will generally be affected by motion in the
3L-1 spectator modes of vibration. The motion in these
other modes reduces the Rabi frequency in much the
same way as lattice vibrations affect a single emitter or
scatterer embedded in a crystal, as described by the
Debye-Waller effect [279,280]. From Eq. (36), we have
the Rabi frequency for the jth ion (assuming all modes
are in specific Fock states)
V
j
n
k
',n
k
=
V
j(o)
n
k
',n
k
U
k{n
pÞk
}
P
pÞk
e
i
h
j
p
(a
p
+a
p
)
{n
pÞk
}l
U
(122)
=
V
j(o)
n
k
',n
k
P
pÞk
e
21/2(
h
j
p
)
2
+
n
p
((
h
j
p
)
2
),
where
V
j(o)
n
k'
,n
k
V
kn'
k
exp(i
h
j
k
(a
k
+ a
k
))n
k
lis the Rabi
frequency of the kth mode (selected for logic), in the
absence of the 3L-1 spectator modes labeled by index p
and vibrational number n
p
. We assume the vibrational
states of spectator modes are independently thermally
distributed with mean vibrational number n¯
p
:
P
n
p
=
(n¯
p
)
n
p
(1 + n¯
p
)
n
p
+1
. (123)
If we average over this distribution, the mean and mean-
squared Rabi frequencies are given by
V
j
n
k
',n
k
=
V
j(0)
n
k
',n
k
P
pÞk
FO
`
n
p
=1
P
n
p
e
21/2(
h
j
p
)
2
+
n
p
S
(
h
j
p
)
2
DG
=
V
j(0)
n
k
',n
k
P
pÞk
e
2(
h
j
p
)
2
(n¯
p
+1/2)
(124)
(
V
j
n
k
',n
k
)
2
=(
V
j(0)
n
k
',n
k
)
2
P
pÞk
FO
`
n
p
=1
P
n
p
e
2(
h
j
p
)
2
+
n
p
S
(
h
j
p
)
2
D
2
G
=(
V
j(0)
n
k
',n
k
)
2
P
pÞk
e
22(
h
j
p
)
2
(n¯
p
+1/2)
I
0
S
2(
h
j
p
)
2
Ïn¯
p
(n¯
p
+1)
D
where we have used the Laguerre polynomial sum iden-
tities
o
`
n=0
z
n
+
n
(x)=
e
2
zx
12z
1 2 z
,
o
`
n=0
z
n
+
n
(x)
2
=
e
2
2zx
12z
1 2 z
I
0
S
2xÏz
1 2 z
D
(125)
In these expressions, z< 1 and I
0
(
e
) is the zeroth
modified Bessel function with argument
e
.
The first expression in Eq. (124) shows the exponen-
tial reduction of the Rabi frequency due to motion in
modes p (Debye-Waller factor). For each mode p,we
can write (
h
j
p
)
2
(n¯
p
+
1
/2)=k
1
/2k
2
eff
(x
j
p
)
2
l where k
eff
is the
effective k vector of the radiation and x
j
p
is the amplitude
of the component of motion parallel to k
eff
for the jth
ion. For L large, the product term on the right side of the
first expression is approximately equal to exp(2
1
/
2 k
2
eff
(x
j
rms
)
2
) where x
j
rms
is the total rms amplitude of mo-
tion parallel to k
eff
. This is the Debye-Waller reduction
factor due to the thermal energy of the jth ion.
The second line in Eq. (124) allows us to determine
the fractional fluctuations in the Rabi frequency from
experiment to experiment
D
V
j(rms)
n
k
',n
k
V
j
n
k
',n
k
Î
S
V
j
n
k
',n
k
D
2
2
S
V
j
n
k
',n
k
D
2
V
j
n
k
',n
k
= Î
F
P
pÞk
I
0
S
2(
h
j
p
)
2
Ïn¯
p
(n¯
p
+1)
DG
2 1
Î
O
pÞk
(
h
j
p
)
4
n¯
p
(n¯
p
+ 1) . (126)
In this last approximation, the Bessel function is ex-
panded to lowest order [I
0
(
e
)=1+
e
2
/4 + ???] which is
appropriate if the 3L-1 arguments of the Bessel function
in Eq. (126) are all small compared to 1. This is ex-
pected to be the case if all modes are cooled to the
Lamb-Dicke limit [(
h
j
p
)
2
n¯
p
<< 1]. Eq. (126) describes
the fractional rms fluctuations in the Rabi frequency
due to thermal motion in the spectator modes of vibra-
tion.
Typically, the thermal motion is determined by initial
conditions and the reservoir is “turned off” once the
experiment begins. This is expected to be the case for
imperfect initial laser-cooling, resulting in a probability
distribution of stable Rabi frequencies with the above
mean and rms values. In this case, the Rabi frequencies
for each ion maintain a constant value during a single
run of the experiment, but deviate from the mean as
304
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
indicated in Eq. (126). Stating the results in another way,
if the total number of modes is large, then the distribu-
tion of Rabi frequencies is nearly Gaussian, and the
probability that a given run of the experiment results in
a Rabi frequency which is fractionally within
e
of the
mean Rabi frequency
V
n
k
',n
k
is
Pr
1)
V
j
n
k
',n
k
2
V
j
n
k
',n
k
V
j
n
k
',n
k
)
<
e
2
erf
1
e
Î2
O
pÞk
(
h
j
p
)
4
n¯
p
(n¯
p
+1)
2
. (127)
For example, if each of 100 spectator modes is laser-
cooled to n¯
p
= 0.1 and has Lamb-Dicke parameter
h
j
p
= 0.01, then the probability that the Rabi frequency
is within 10
24
of the mean value is approximately 0.23.
To see how Eq. (126) scales with the number of ions,
assume that k
eff
is aligned with the axis of a linear trap,
where the axial motion is described by L modes. If the
frequencies and amplitudes of all modes contributing to
the axial motion of ion j are assumed to be about the
same, we can write
h
j
p
.
h
1
/L
1/2
and n¯
j
p
. n¯ where
h
1
and n¯ are the Lamb-Dicke parameter and mean occupa-
tion for the axial motion of a single (thermalized)
trapped ion. In this case, Eq. (126) becomes
D
V
j(rms)
n
k
',n
k
V
j
n
k
',n
k
.
h
2
1
F
n¯(n¯+1)
L
G
1/2
. (128)
This expression indicates that a large number of ions is
beneficial because it tends to average out the effects of
motion in the (L 2 1) extraneous modes. Equation (128)
is an overestimate of the fluctuations since the L 2 1
extraneous modes will have higher frequency than the
COM mode, leading to smaller amplitudes of motion
than assumed in this crude estimate.
The Rabi frequency fluctuations discussed here trans-
late to fractional phase offsets [(
z
k
/
Q
k
of Eq. (94)] for
each gate operation as indicated in Eqs. (124) and (126).
In practice, the mean Rabi frequency
V
n
k
',n
k
can be mea-
sured by averaging over many experiments. If we as-
sume the COM mode is used for logic, then
h
k
=
h
1
/L
1/2
.
Therefore, for L >> 1, we cannot choose
h
1
to be too
small, or else the entangling operations for quantum
logic become too slow (Eq. (24)). To suppress the ef-
fects of Debye-Waller factor fluctuations, it is therefore
desirable to cool all modes (whose motion is parallel to
k
eff
) to the zero-point state.
4.4.6 Coupling to Spectator Levels
We have assumed that each time an external field is
applied to form part of a logic operation, only two quan-
tum states take part in the interaction. This assumption
is valid when any other state is far from resonance. To
explain this in a simple example, we refer to Fig. 9.
Suppose we want to drive a transition between states ↓l
to ↑l with radiation resonant with this transition. If
level ↓l is coupled to level sl with this radiation, the
wavefunction will have a small admixture of state sl
after the operation; the amplitude of state sl will be-
come larger as
D
becomes smaller. After a sequence of
many operations, the amplitudes of these “spectator”
states can build up and cause errors in the computation.
Fig. 9. Simplified energy level diagram for characterization of cou-
pling to, and spontaneous emission from, off resonant “spectator”
levels. We assume coherent radiation is tuned near the transition fre-
quency for the ↓lnl →ln'l transition (for simplicity we have
not shown the motional substructure for the ↓l and ↑l states). The
state sl is assumed to be the nearest spectator state from which
off-resonant coupling can occur.
4.4.6.1 Polarization Discrimination of Internal
States
In an array of trapped ions, two internal states of each
ion comprise a qubit of information. In addition, a third
auxiliary internal level auxl may be required (tran-
siently) for the operation of a CN gate as described in
Sec. 3.3. This state might be a particular Zeeman sub-
level of a hyperfine multiplet [17] or an optical
metastable state [1]. By employing suitable polarizations
of the driving field, particular internal state transitions
can be selected with high discrimination. Furthermore,
the Zeeman sublevels can be spectrally resolved (assum-
ing the Zeeman splitting is much larger than the Rabi
305
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
frequency), by applying a magnetic field. This combina-
tion will help to isolate the two internal levels of interest
from other internal levels. Here, we give an example of
how polarization selection can be used to suppress cou-
pling to internal “spectator” states.
We consider the case of Ref. [17], which uses two 2s
2
S
1/2
ground state F, m
F
l hyperfine levels of
9
Be
+
ions
as qubit levels. These levels are designated ↓l ≡2,2l
and ↑l ≡1,1l (Fig. 5). We also make the identification
auxl ≡2,0l. They are driven using stimulated-Raman
transitions by detuning the lasers from the
2
P
1/2
state. By
choosing the Raman beam coupling between the
2
P
1/2
and ↓l states to be
s
2
polarized and the coupling be-
tween the
2
P
1/2
and ↑l states to be p polarized, the ↓l
and ↑l states form a closed family; that is, neither the
↓l or ↑l state can be driven to other Zeeman sublev-
els. In this case, the nearest off-resonant transitions to
consider are detuned by at least the hyperfine frequency
(. 1.25 GHz in
9
Be
+
). Similarly, by choosing the beam
coupling between the
2
P
1/2
and auxl states to be
s
+
polarized (and the coupling between the
2
P
1/2
and ↑l
states to be p polarized), the auxl level is coupled only
to the ↑l level.
4.4.6.2 Spectral Discrimination of States
Spectator states can include both motional and inter-
nal states. For example, a single CN gate uses pulses
which drive on the carrier as well as first motional
sidebands of the COM mode. Discrimination between
these two types of transitions cannot be made with
polarization selection since the relative strengths of the
matrix elements are independent of polarization. Since,
in the Lamb-Dicke limit, the carrier operation will have
a higher resonant Rabi frequency than the sideband op-
erations, the largest source of contamination may be due
to an off-resonant excitation of the carrier during the
sideband operations. The basic problem can be illus-
trated by considering the simple example illustrated in
Fig. 9. Here, we assume decay from the excited states is
negligible,
G
s
=
G
= 0. We assume we want to carry out
an operation which coherently drives the ↓l ↑l
transition (resonance frequency
v
0
) and avoid the pres-
ence of any amplitude in the “spectator” state sl after
the operation. We assume radiation of frequency
v
.
v
0
is applied which couples levels ↓l and ↑l with
strength
V
. We assume this radiation also couples levels
↓l and sl with (resonant) strength
V
', but, for sim-
plicity, does not couple levels ↑l and sl. In the context
of quantum logic,
V
might correspond to the Rabi fre-
quency for a sideband transition and
V
' the Rabi fre-
quency for the carrier transition. If we take the zero of
energy in Fig. 9 to be the energy of the ↓l level, we can
write the Hamiltonian as H = H
0
+ H
I
, where
H
0
=
"v
0
↑lk↑ +
"
(
v
0
+
D
)slks,
(129)
H
I
=2
"
cos
v
t(
V
[↓lk↑ + ↑lk↓]
+
V
'[↓lks+ slk↓]),
where the expression for H
I
is equivalent to that in Eq.
(14). In the Schro¨dinger picture, we write
C
= C
↓l +
C
exp(2 i
v
0
t)↑l + C
s
exp(2 i[
v
0
+
D
]t)sl. This leads
to equations for the amplitudes
C
˙
= 2 i
V
e
i
d
t
C
2 i
V
'e
i(
d
2
D
)t
C
s
,
C
˙
= 2 i
V
e
2i
d
t
C
, (130)
C
˙
s
= 2 i
V
'e
2i(
d
2
D
)t
C
,
where
d
v
2
v
0
and we have neglected rapidly vary-
ing terms proportional to exp(6 i(
v
+
v
0
)t) (the usual
rotating wave approximation). If we make the substitu-
tion C
s
= C
'
s
e
i
D
t
, the equation for C
˙
s
can be written
C
˙
'
s
+ iDC
'
s
= 2 i
V
'e
2idt
C
. (131)
For
V
' small enough relative to
D
,wecan
“adiabatically eliminate” level s, by assuming dC
s
'/
dt<< DC
'
s
. In this case, Eqs. (130) become
C
˙
. 2 i
V
e
idt
C
+ i
(
V
')
2
D
C
,
C
˙
= 2 i
V
e
2idt
C
, C
'
s
. 2
V
'
D
e
2idt
C
. (132)
The Stark shift term
d
s
(
V
')
2
/
D
is a downward energy
shift of the ↓l state. It can be suppressed in the previ-
ous equation by including it into a shift of
v
0
. That is, by
making the substitution C
= C
'
exp(i
d
s
t) and choosing
v
2
v
0
=
d
s
,wefind
C
˙
'
= 2 i
V
C
, C
˙
= 2 i
V
C
'
, C
'
s
= 2
V
'
D
C
'
. (133)
In this case, C
'
and C
are given by Eqs. (17), leading
to a the desired evolution between the ↓l and ↑l
states. Moreover, if the couplings
V
(and
V
') are turned
on and off slowly compared to 1/
D
(the condition of
adiabaticity), then the amplitude in state sl adiabati-
cally grows during the operation and reduces to zero
upon completion of the operation, since it is propor-
tional to
V
'.
The general problem, however, is that we desire to
make all transitions as fast as possible. This means that
we cannot satisfy the adiabatic condition; therefore C
'
s
306
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
will not be simply proportional to
V
' and will have some
nonzero value after
V
' is reduced to zero. For example,
if
V
' is turned on adiabatically, but turned off diabati-
cally, C
s
will have a final amplitude with magnitude
V
'C
'
/
D
. Estimates of these effects in specific con-
texts are given by Poyatos et al. [261] and James [61].
Since it should be possible to separate internal states
by relatively large frequencies and further discriminate
from internal spectator states with polarization, the most
important task will be to suppress transitions to mo-
tional spectator levels. The fundamental problem is that
in order to discriminate between carrier and sideband
transitions, the ion must be able to tell that it is moving.
If the time of the operation
t
op
is small compared to the
inverse of the frequency of the motional mode, this is
impossible. Therefore, the fundamental limit on
t
op
is
given approximately by 1/
v
COM
. To maximize computa-
tional speed, we desire to approach this condition as
closely as possible. This can be aided by using Rabi
envelopes which can suppress certain spectral compo-
nents [281,282]. Since the general problem will involve
more than one spectator level (more motional states), a
more general approach might be to apply optimal con-
trol theory [283–286].
4.4.6.3 Tailoring of Laser Fields
As discussed in the last section, the speed of an ion-
trap quantum computer will be fundamentally limited
by excitation of motional sidebands other than the de-
sired one. To see this in a specific case, the conditional
dynamics for the first CN gate described in Sec. 3.3
occurs in step (1b). That step uses a 2p pulse on 1st blue
sideband of the ↑l1l ↔auxl0l transition. To make
this step fast, we would like to increase the field inten-
sity and decrease the interaction time accordingly. How-
ever, as the pulse becomes shorter, the spectrum of the
pulse becomes wider. Eventually, the spectral width of
this pulse is larger than
v
z
so that we drive the relatively
strong and unwanted ↑l1l auxl1l carrier transi-
tion. In general, we also drive ↑l1l auxln'l tran-
sitions, where n' > 1. Most of these unwanted transitions
can be suppressed by making all terms in the expansion
in of Eq. (28) equal to zero except the term proportional
to B/z (or E/z for electric dipole transitions). In this
case, we suppress coupling to all unwanted auxl levels
except the auxl2l state whose resonant frequency is
detuned from the ↑l1l auxl0l transition by 2
v
z
.
Suppression of even or odd order terms in Eq. (28)
can be accomplished by appropriate positioning of
standing waves relative to the ion [40,61,287,288]. To
illustrate with an example, we first examine the transi-
tion ↓l0l ↑l1l. For simplicity, we assume single
photon electric-dipole (rather than stimulated-Raman)
transitions and consider only motion along the z direc-
tion. We want to suppress all terms in the expansion of
the field [Eq. (28)] except the E/z term. We need to
synthesize, using M separate standing wave laser beams,
a wave which has the form
E = E
0
FO
M
m=1
C
m
sin(k
m
z)
G
cos(
v
t), (134)
where k
m
k
m
? zˆ, and where z = 0 corresponds to the
equilibrium position of the ion in question. The case of
M = 1 has already been suggested in various contexts to
suppress coupling to the carrier (Dn = 0) transition (see
Ref. [40] and Sec. 4.4.4 above). For M >1,wewantto
satisfy
C
1
k
n
1
+ C
2
k
n
2
+ ... + C
M
k
n
M
=0, n = 3,5, ... 2M 21. (135)
In this case, the next sideband spectator level which has
nonzero coupling is one where n' 2 n=2M + 1. This
transition is detuned by 2M
v
z
from the transition of
interest and the coupling is suppressed by approximately
h
2M
=(k
1
z
0
)
2M
compared to
V
1,0
or
V
0,1
.
Second, we consider carrier transitions ↓lnl
↑lnl, where n = 0,1. With the stimulated-Raman
technique on hyperfine transitions, transitions of the
type ↓lnl ↑ln'l, n Þ n', are highly suppressed
by the use of copropagating beams. For single photon
transitions, we need to synthesize, using separate laser
beams, a wave which has the form
E = E
0
FO
M
m=1
C
m
cos(k
m
z)
G
cos(
v
t). (136)
For M >1,wewant
C
1
k
n
1
+ C
2
k
n
2
+ ... + C
M
k
n
M
=0,
n = 2,4, ... 2(M 2 1). (137)
In this case, the next spectator level which has nonzero
coupling is one where n' 2 n=2M. This transition is
detuned by 2M
v
z
from the carrier and the coupling is
suppressed by approximately
h
2M
=(k
1
z
0
)
2M
compared
to
V
1,1
or
V
0,0
. For the carrier transitions, we do not have
the problem with the remaining spectator level as noted
for the first sideband case above. On the other hand, for
logic operations which use the carrier transition (Sec.
2.3 and Ref. [174]), we do not want
h
to be too small,
or else the gates take to long to implement. Therefore
the suppression of the higher-order sidebands may not
be as great as in the case of logic using first sidebands.
For qubits coupled by single-photon transitions, it
may be difficult to suppress couplings to motional
307
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
modes in the x and y directions. This will be true if we
carry out logic on a string of ions in a linear trap and use
the axial COM mode. The laser beams must be at an
angle with respect to the z axis to be able to select
particular ions. This kind of problem can be eliminated
by using the x or y COM mode for logic and making k
orthogonal to the other directions. Finally, we remark
that the use of standing waves will give rise to optical
dipole potentials which can shift the ion motional fre-
quencies. These shifts should be small, and they can be
incorporated into the definitions of the motional oscilla-
tion frequencies.
4.4.6.4 Spontaneous Emission
The internal atomic states of trapped ions, which
store quantum bits of information, must be protected
from spontaneous emission, at least for the duration of
the computation. This excludes the possibility of “error
correction,” (See Sec. 3.3) which can tolerate a certain
level of errors due to spontaneous emission. As dis-
cussed in Sec. 4.2.1, for qubit levels coupled by single
photon optical transitions, this may be accomplished by
employing long-lived energy levels which do not have an
allowed electric dipole coupling, such as metastable
electronic levels with a quadrupole or intercombination
coupling to the ground state. Although the interaction of
these states with the vacuum (causing spontaneous
emission) is reduced, their interaction with an external
field for use in quantum logic operations is also reduced.
This results in a fundamental limit on the accuracy of
each operation by roughly the ratio of the spontaneous
emission rate to the Rabi frequency
j
=
G
/
V
. In the case
of optical transitions,
V
cannot be increased indefi-
nitely, since at optical intensities beyond about 10
14
W/
cm
2
, the atom is quickly photoionized. This amounts to
inaccuracies due to spontaneous emission on the order
of
j
=10
26
2 10
27
[289]. Even this limit may be too
optimistic, as the two-level approximation breaks down
before photoionization occurs, and the coupling to other
electronic levels must also be considered [188,289].
This results in inaccuracies due to spontaneous emission
on the order of
j
=10
25
2 10
26
, depending on the par-
ticular ion used. To understand the nature of the prob-
lem, we examine a simplified system. More detailed
treatments are given in Refs. [188] and [289].
Consider the situation shown in Fig. 9. We assume
that levels ↓l and ↑l comprise the qubit states. We
want to drive coherent transitions between these two
levels but, in general, we must consider spontaneous
Raman scattering from other non resonant spectator
levels. For this approximate treatment, we will consider
only the effects of the nearest, most strongly coupled
state which is designated as state sl in the figure. For
simplicity, we assume that levels sl and ↑l are cou-
pled only to the ground state ↓l (and not each other) by
the applied radiation, and that they decay by the same
coupling process to the ground state with rates
G
s
and
G
. To allow for different coupling processes for the sl
and ↑l states, we let
G
=
kG
s
and we neglect
v
3
factors
in the differences of the lifetimes (this would be valid if
the ↓l →l transition frequency is much larger than
D
). The worst case is given by considering sideband
transitions between levels ↓l and ↑l. From Sec. 2.3.1,
the resonant Rabi frequency is given by
V
1
.
hV
,
.
hk
1/2
V
s,
. We assume that during the operation, the ion
has a probability of about 0.5 of being in the ↓l and
↑l state. With these assumptions, the ratio
j
of sponta-
neous emission (from both upper levels) to Rabi fre-
quency is
j
=
R
s
V
1
.
1
2
G
+
1
2
G
s
S
V
s,
D
D
2
=
G
s
2
V
1
S
k
+
z
k
D
, (138)
where
z
V
2
1
/(
hD
)
2
, and the factors of
1
/2 come from the
probabilities the ion is in the ↓l or ↑l state. The most
optimistic answer is given by minimizing
j
with respect
to
k
which leads to
j
.
G
s
/(
hD
). As an approximate
“best case” we take
G
s
/2p = 25 MHz,
D
/2p =10
14
Hz,
V
1
/2p = 10 MHz,
h
= 0.1, giving
j
. 2.5 3 10
26
and
requiring
k
. 0.4
j
and
G
. 10
26
G
s
. Therefore, weakly
allowed transitions are desirable if single photon optical
transitions are used for qubit transitions.
In the case of two-photon stimulated-Raman transi-
tions between stable electronic ground states, the ratio of
spontaneous emission rate to Rabi frequency is approxi-
mately
j
SR
=
g
se
/(g
2
/
D
R
), where
g
se
G
g
2
/
D
2
R
is the off-
resonant spontaneous emission rate, g is the resonant
single-photon Rabi frequency of each laser beam, and
D
R
is the detuning of the Raman beams from the excited
state (Sec. 2.3.3). This results in an inaccuracy
G
/
D
R
due
to spontaneous emission, which is independent of opti-
cal intensity. Since Raman transitions between S elec-
tronic ground states are effective only when the detun-
ing
D
R
is not much greater than the fine structure
splitting of the atom [290]; this results in an inaccuracy
j
SR
due to spontaneous emission in range from about
10
24
(
9
Be
+
)to10
27
(
199
Hg
+
), depending on the particular
ion used. Spontaneous emission from spectator elec-
tronic levels should not significantly affect this limit,
provided that their splitting from the virtual excited
state significantly exceeds
D
R
and that the single photon
resonant Rabi frequencies coupling the ground states to
the spectator levels are not much bigger than g [188].
These appear to be reasonable assumptions for most
candidate ions.
308
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
The decohering effects of spontaneous emission can
be overcome by error correction schemes. Error correc-
tion is complicated by the fact that when spontaneous
emission occurs, the atoms may decay to states outside
the original set of computational basis states. However,
this situation can, in principle, be detected by optically
pumping the ions back to the computational basis and
applying the error correction schemes [291,292].
Spontaneous emission decoherence could, in princi-
ple, be nearly eliminated by driving single-photon tran-
sitions between ground-state-hyperfine or Zeeman lev-
els with rf or microwave radiation since spontaneous
emission from these levels is negligible. Refs. [3] and
[293] discuss this possibility, where inhomogeneous
magnetic fields couple the internal and motional states.
The speed of sideband operations is limited by the size
of the field inhomogeneity one can obtain. From Eq.
(28), we want a coupling Hamiltonian H
I
= 2
m
x
(B
x
/
z)z. Assume B
x
/z =(B
x
/z)
0
cos
v
t,
m
x
=
m
M
S
x
,
and z = z
0
(a + a
) as in Sec. 2.3.1. For resonance at the
first red sideband (
v
=
v
0
2
v
z
), H
I
is given by Eq. (27)
with
hV
= 2
m
M
(B
x
/z)
0
z
0
/(4
"
). If we take
m
M
=
m
B
,
z
0
= 10 nm, then to achieve
V
1
/2p = 1 MHz, we require
(B
x
/z)
0
. 290 T cm
21
; a difficult task. Moreover, it
would be difficult to address selected ions because of
the long wavelength of the radiation relative to typical
ion spacings (see also Sec. 5.3).
4.4.7 Mode Cross Coupling During Logic
Operations
In the preceding sections, we have assumed that when
transitions are driven between ↓l
j
n
k
l and ↑l
j
n
'
k
l
involving a single mode of motion k, the other 3L-1
spectator modes of motion are not affected because
coupling to them is nonresonant. However, when the
sum or difference frequency of two or more spectator
modes is near the frequency of the desired mode-k tran-
sition (
v
k
n
'
k
2 n
k
), higher order couplings can entan-
gle the ↓l
j
n
k
l and ↑l
j
n
'
k
l states with the spectator
mode states.
Equation (33) describes the general interaction
Hamiltonian between the internal levels of ion j and all
3L modes of motion. By expanding the exponential in
Eq. (33) to all orders, we find
H'
Ij
=
"V
'
F
S
+j
SP
3L
l=1
O
`
b
l
,d
l
=0
(i
h
j
l
a
l
)
b
l
(i
h
j
l
a
l
)
d
l
b
l
!d
l
!
3 e
i(b
l
2d
l
)
v
l
t
D
e
2i(
d
t2
f
j
)
+ h.c.
G
, (139)
where
V
' =
V
exp[2
1
/2
S
i
(
h
j
i
)
2
]. This equation describes
the processes of each mode l gaining or losing (b
l
2 d
l
)
vibrational quanta, accompanied by the raising or lower-
ing of the internal electronic levels of ion j. In general,
we must account for all terms in Eq. (139) which do not
vary rapidly in time, or terms in which the resonance
condition is nearly met: S
l
(b
l
2 d
l
)
v
l
.
d
=
v
k
(n
k
' 2
n
k
). Although detailed treatment of this problem is be-
yond our intent, a couple of comments may be made.
In general, we must account for all the terms in Eq.
(139) which cause significant errors in the overall com-
putation we are trying to carry out. For two or more
trapped ions, some combination of modes will nearly
always satisfy the resonance condition. However, this
may occur only for high orders of b
l
and d
l
, and if the
Lamb-Dicke criterion is met, the contributions are van-
ishingly small. The terms that will cause problems are
the ones that are close to satisfying the resonance condi-
tion and are relatively low order in b
l
and d
l
.Ifthe
Lamb-Dicke criterion is satisfied, it will always be pos-
sible to avoid these spurious couplings, but it may be at
the expense of making the Rabi frequency so small (in
order to avoid coupling to relatively nearby off-resonant
terms) that the operations will become too slow.
To understand this problem in the context of a simple
example, we assume that a cross-mode coupling of this
type occurs when two modes, p and q, have frequencies
which satisfy the condition n
p
v
p
2 n
q
v
q
. 0, or 6
v
k
corresponding to possible extraneous mode coupling on
the carrier, first blue sideband, or first red sideband of
the logic operations (assumed to utilize mode k). This
additional resonance condition yields, to lowest order in
the Lamb-Dicke parameters, the effective Hamiltonian
H
Ij
'
=
"V
'S
+j
H
1+i
h
j
k
(a
k
e
2
v
k
t
+ a
k
e
i
v
k
t
)
+
(i
h
p
a
p
)
n
p
(i
h
q
a
q
)
n
q
n
p
!n
q
!
e
i(n
p
v
p
2n
q
v
q
)t
J
e
2i(
d
t2
f
j
)
+ h.c., (140)
where the resonance conditions are given by
d
v
2
v
0
=0,+
v
k
,or2
v
k
. A specific example is relevant to
the NIST single
9
Be
+
ion experiments. Here, mode k was
the x oscillation, and modes p and q are identified with
the z and y oscillations of the single ion in the trap. In
this experiment
h
x
= k ? xˆx
0
,
h
y
= k ? yˆy
0
,
h
z
= k ? zˆz
0
,
and
v
x
.
v
z
2
v
y
. (The frequency relationship
v
x
=
v
z
2
v
y
is a consequence of Maxwell’s equations
for a quadrupole rf trap in the absence of static poten-
tials applied to the electrodes [211].) We assume that the
desired transition is the first blue sideband of mode
x(
d
=
v
x
). In this case, the resonant part of Eq. (140)
becomes
H
I
'
.
"V
'
F
S
+
H
i
h
x
a
x
2
h
z
h
y
a
z
a
y
+ O(
h
3
)
J
+ h.c.
G
.
(141)
309
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
The term proportional to a
x
is the desired anti-Jaynes-
Cummings operator, and the term proportional to a
z
a
y
can entangle the internal state with the other spectator
modes (z and y), leading to errors.
For logic operations on a string of ions in a linear
trap, we will assume that all other mode frequencies are
higher. With the use of stimulated-Raman transitions,
we can make Dkizˆ and restrict our attention to spectator
modes along the z axis. Nevertheless, as L becomes
large, nearby resonances of the type shown in Eq. (140)
will become harder to avoid. These coupling terms al-
ways scale as products of Lamb-Dicke parameters. Thus
if the spectator mode Lamb-Dicke parameters are small
enough, or if at least one Lamb-Dicke parameter is ap-
proximately zero, the higher order unwanted resonances
may be sufficiently suppressed. Furthermore, if the
spectator modes are cooled to near the zero-point energy
(knˆl << 1), then any couplings in Eq. (140) with powers
of the annihilation operator a
q
will be absent most of the
time. Hence, in large registers, it will probably be im-
portant to cool all modes to near the zero-point energy.
5. Variations
5.1 Few-Ion Accumulators
Many of the problems anticipated above, such as un-
wanted coupling to adjacent ions or spectator modes,
will be aggravated by a large number of ion qubits in a
single ion trap “register.” Moreover, implementation of
quantum error correction schemes will require highly
parallel processing [294]. Therefore, a multiplexing
scheme for ion qubit registers is desirable. One possibil-
ity is to perform all logic in minimal accumulators
which hold one or two ions at a time [65]. Ions would be
shuffled around in a “super-register” and into and out of
(multiple) accumulators which are well shielded from
the other ions. The shuffling could be accomplished
with interconnected linear traps with segmented elec-
trodes as shown schematically in Fig. 10; construction
of such traps appears to be possible with the use of
lithographic techniques [295]. Single-bit rotations on
the mth ion would be accomplished by moving that ion
into an accumulator. Logic operations between ions m
and k would be accomplished by first moving these ions
into an accumulator. An accumulator could also hold a
second species of ion (say Mg
+
) which could be used to
provide laser cooling to the n =0l state of the motional
mode used for the logic operations, if necessary. There-
fore, for logic operations, an accumulator would hold
two computational ions and the auxiliary ion. This
scheme should make it easier to select ions with laser
beams because it should be possible to address one ion
while nulling the laser intensity on the other ion, even
with very high trap frequencies (see Sec. 4.4.4). The
very small number of logic ions in an accumulator (1 or
2) would make extraneous mode coupling much easier
to avoid. The main disadvantage appears to be that com-
putational speed is reduced because of the time required
to shuffle ions in and out of the accumulator and provide
laser cooling with the auxiliary ion, if required. How-
ever, energy shifts of the ion’s internal structure, due to
the electric fields required to move the ion, need not be
severe. For example, to move a
9
Be
+
from rest to a
location 1 cm away (and back to rest) in 1 ms would
require a field of less than 50 V/cm. Electric fields of
this order should give negligible phase shifts in qubits
based on hyperfine structure (Sec. 4.2.3). The accompa-
nying phase shift caused by time dilation for a transition
Fig. 10. Schematic diagram of an ion trap “super-register” containing
few-ion accumulators in which logic operations are performed. We
assume the trap is generically equivalent to the linear trap of Fig. 1.
The rods of the trap in Fig. 1 are replaced by segmented electrodes in
which adjacent segments are at the same rf potential but where the
segments support different static potentials. This allows ions to be
selectively moved around an in and out of ion accumulators. In the
accumulator shown a logic operation between ions k and m is indi-
cated. The configuration of the laser beams is chosen to null the
intensity on ion m while performing an operation on ion k. The beams
must be reconfigured to perform an operation on ion m while leaving
ion k unaffected. The auxiliary ion may be required to perform laser
cooling each time new logic ions are moved into the accumulator.
310
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
frequency
v
0
/2p = 1.25 GHz (hyperfine structure in
9
Be
+
) would be less than 1 mrad.
5.2 Multiplexing With Internal States
In principle, it should be possible to multiplex quan-
tum information into all 3L modes of ion motion. This
will probably be useful in experiments which use a
small number of ions. However, because of the higher
potential for decoherence of motional modes, it might
prove more useful to multiplex quantum information in
multiple internal states. This would give more work
space with a smaller number of ions. Quantum logic
within multiple internal states would most likely follow
the ideas of NMR quantum computing [165,166]. A
simple example is given by the eight ground state hyper-
fine levels in
9
Be
+
which can be labeled like states of
three coupled spin-
1
/2 particles
e
1
,
e
2
,
e
3
l where
e
i
[ {0,1}. For instance, a Toffoli gate [140] is realized
by driving a p pulse on the F,M
F
l = 2,2l →1,1l
transition in the ground state hyperfine manifold of
9
Be
+
if we label the states 2,2l ≡1,1,0l and 1,1l ≡1,1,1l,
in the notation F,M
F
l ≡
e
1
,
e
2
,
e
3
l. Technically, the
readout of specific internal quantum states would be
more complicated since the methods employed in the
experiments so far distinguish between only two possi-
bilities, seeing fluorescence or not seeing fluorescence,
on any given ion. However, by appropriate mappings,
superpositions of any two internal states could be
mapped onto two states of “readout” ions and the detec-
tion accomplished as described above. To multiplex in-
formation into internal states, we need a way to map
qubits from any ion into the internal levels of other ions.
To give a specific example of how this can be accom-
plished, we consider the case of
9
Be
+
where we might
wish to multiplex into the F,M
F
l = 2,2l, 1,1l, 2,1l,
and 1,0l states shown in Fig. 11. This figure is essen-
tially the same as Fig. 5, but for simplicity of notation,
we label these hyperfine states as 0l, 1l, 0'l,and1'l
respectively. We want to accomplish the transformation
C
i
=
a
u0l
A
u0l
B
+
b
u0l
A
u1l
B
+
g
u1l
A
u0l
B
+du1l
A
u1l
B
S
a
u0l
A
+
b
u1l
A
+
g
u0'l
A
+du1'l
A
D
u0l
B
,
(142)
between ions A and B. This can be realized with the
transformation
M
ˆ
= C
ˆ
A',B
? C
ˆ
A,B
? C
ˆ
B,A'
? C
ˆ
B,A
?P
ˆ
A,1,0'
, (143)
where the operator P
ˆ
A,1,0'
denotes a p transition between
the states u1l
A
and u0'l
A
on ion A and the operators C
ˆ
c,t
are defined in Sec. 3.3. For example, C
ˆ
A',B
denotes a
controlled-not operation with the u0'l
A
and u1'l
A
states
acting as the control bit states and the u0l
B
and u1l
B
states
acting as the target bit states.
5.3. High-Z Hyperfine Transitions
The possibility of using highly charged ions as qubits
poses some interesting possibilities. For simplicity, we
consider using hyperfine states in high-Z hydrogenic
ions as qubit levels where Z is the nuclear charge.
More complicated high-Z ions appear to show similar
features. If Z is high enough, hyperfine transitions occur
in the optical region of the spectrum. Therefore, the
required high field gradients necessary for quantum
logic would be provided by the laser fields used to drive
the hyperfine transitions. The potential advantages
are: (1) Off-resonant spontaneous scattering, which is a
source of decoherence in other cases (Refs. [188], [289],
and Sec. 4.4.6.4), is essentially eliminated since the first
excited 2p level is at much higher energy above the
ground state. (2) The trap binding of high-Z hydrogenic
ions can typically be stronger that singly-ionized atoms,
thereby increasing motional frequencies and reducing
the time of fundamental operations. The disadvantages
are (1) High-Z ions are hard to produce. (2) The life-
times of the upper hyperfine states are short enough that
Fig. 11. Hyperfine levels of the 2s
2
S
1/2
ground state of
9
Be
+
in a weak
magnetic field (not to scale). The energy levels are designated by
horizontal lines. Above the lines, the levels are represented by atomic
physics labels (F,M
F
) where F is the total angular momentum and M
F
is the projection of the angular momentum along the magnetic field
axis. The separation of Zeeman substates in the different F manifolds
is approximately equal to 0.7 3 10
10
B
0
Hz where B
0
is expressed in
teslas. The separation of the F = 1 and F = 2 manifolds is approxi-
mately 1.25 GHz at B
0
= 0. For simplicity of notation, in Sec. 5.2, we
make the identifications F,M
F
l = 2,2l ≡0l, 1,1l ≡1l, 2,1l
0'l, 1,0l ≡1'l.
311
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
spontaneous emission decoherence from these states
cannot be neglected. (3) Detection of internal states may
be difficult.
The ion production problem is not intractable; for
example, recent storage ring measurements have been
able to determine the transition wavelengths and life-
times of hydrogenic
209
Bi
82+
[296] and
165
Ho
66+
[297].
Effort in some laboratories is directed to transferring
similar high-Z ions to low energy ion traps. The hyper-
fine frequency of a hydrogenic ion can be estimated to
be
v
hfs
/2p 2.54310
8
Z
3
Fug
n
u Hz, where F and F-1 are
the values of angular momentum in the hyperfine states
and g
n
is the nuclear g-factor (
m
=
m
n
Ig
n
, m
n
= nuclear
magneton, I = nuclear spin) [298]. The spontaneous
decay rate from the excited state is estimated to be
g
rad
=1/
t
rad
1.01310
–42
[M
ab
/(2F
b
+1)] (
v
hfs
/2p)
3
s
–1
,
where the term in square brackets [298] depends on
matrix elements and is on the order of 1. For example,
the hyperfine transition in
209
Bi
82+
has a wavelength of
244 nm and the upper level decays with a lifetime of
351 ms [296]. For
165
Ho
66+
, the transition wavelength is
573 nm and the decay time is 2 to 3 ms [297]. These
experimental numbers agree roughly with the formulas
above. These wavelengths are reasonable for quantum
logic, but the lifetimes are somewhat short. However
since the lifetime scales as 1/Z
9
, it might be possible to
use lower Z, longer wavelength transitions. The readout
of the qubits appears difficult, because coupling to a
higher electronic level and observing fluorescence is
impractical due to the short wavelengths involved. In
principle, however, the qubit information from these
ions could be transferred to other ions whose states are
easier to read, perhaps using small accumulators as
described in the previous section and in Sec. 4.3.
6. Other Applications
6.1. Quantum Correlations
In a classic Paper, Einstein, Podolski, and Rosen
(EPR)[299] argued that quantum mechanics provides an
incomplete description of physical reality and specu-
lated that it might be an approximation to some theory
which would provide a complete description. Today
such theories, which supplement quantum mechanics
with additional, as-yet unobserved parameters, are
called hidden-variables theories. If such a theory obeys
some physically reasonable conditions forbidding action
at a distance, it is called a “local hidden-variables
theory.”
Simple quantum logic gates performed on small
numbers of trapped ions can lead to interesting experi-
ments which may shed light on the viability of local
hidden-variables theories. For example, as described by
Cirac and Zoller [1] and Sec. 3.4, using controlled-not
gates, we can generate the state
C
=
1
Ï2
(ul
1
ul
2
+e
i
f
ul
1
ul
2
). (144)
In the context of EPR, this is called a Bell state. Accord-
ing to Bell’s theorem [300], such a state allows us to
distinguish between quantum mechanics and all local
hidden-variables theories. If the states of the two en-
tangled particles are detected outside of each other’s
light cones, then, for particular sets of measurements,
we may derive inequalities which all local hidden-
variables theories must obey, but which quantum
mechanics violates. The experiments performed by
Clauser and Shimony [301] and Aspect et al. [302]
provide strong evidence against local hidden-variables
theories. Their work used polarization measurements on
entangled pairs of photons. In their experiments, the
detection of the photons’ polarization states occurred
outside each others’ light cones. Thus, the measurement
on one photon could not possibly have affected the other
measurement, which closed possible “loopholes” in the
proof of quantum mechanics over other explanations.
However, some loopholes still remain open. Since the
photon detection in the Aspect, et al. experiments was
not 100 % efficient, the group had to make assumptions
that the photons they measured were a “fair” sample of
the whole population of events. Thus, their experiments
do not rule out the (seemingly implausible) possibility
of local hidden-variables theories in which the hidden
variables cause some sub-ensemble of the photon
pairs to preferentially interact with the measurement
apparatus.
In the system of two ions, we may detect the state of
either ion with nearly 100 % efficiency through the use
of “electron shelving” (Sec. 2.2.1). On the contrary, it
will be difficult to perform measurements on two ions
outside each other’s light cone. Such a measurement
would require separating the ions by a distance larger
than the speed of light times the measurement time or
transferring quantum information over large distances
[151]. (In principle, the ions could be first entangled and
then placed in different traps which could be separated
by large distances before measurements were per-
formed.) Nonetheless, an experiment with two trapped
ions could be viewed as complementary: the photon
experiments definitively close loopholes of causality,
and the ion experiments could close loopholes due
to detection inefficiency. Such experiments have the
additional appeal of studying EPR on massive particles
[303]. EPR states of atoms have recently been created
in an atomic beam using the methods of cavity QED
[304]; if detection efficiency can be improved, these
experiments could also close loopholes due to detection
inefficiency. Moreover, even though measurements of
quantum correlations between entangled ions cannot be
312
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
easily performed outside each other’s light cone, one
can argue strongly that the ions cannot transfer infor-
mation by any known mechanism. Therefore, if the
observed correlations violate Bell’s inequalities, the
correlations are established by some new force of nature
or are, in fact, inherent in the structure of quantum
mechanics.
An intriguing possibility for ions is the possibility of
making “GHZ states” [54,305,306]. For three ions, the
GHZ state is
C
=
1
Ï2
(ul
1
ul
2
ul
3
+e
i
f
ul
1
ul
2
ul
3
). (145)
For such a state, a single measurement can distinguish
between the predictions of quantum mechanics and
those of any local hidden-variables theory [54,305].
Aside from this, Bell states, GHZ states, and
Schro¨dinger-cat states are highly entangled, and are thus
of inherent interest for the study of uniquely quantum
behavior. As the experiments improve, it will be interest-
ing to push the size of entangled states to be as large as
possible. The question is not whether we can make states
which have the attributes of Schro¨dinger cats, but how
big can we make the cats? Certain theories which
address the measurement problem will be amenable to
experimental tests, for example, quantitative limits on
spontaneous wavefunction collapse theories [307,308]
can be established. The isolation from the environment
exhibited by trapped ions, coupled with the control
possible over their quantum state and high detection
efficiency make them an interesting laboratory for the
study of fundamental issues in quantum mechanics.
6.2. Simulations
The nonlinearities with respect to motional raising
and lowering operators inherent in the coupling
Hamiltonians of Eqs. (14), (37) and (48) can lead to a
rich variety of dynamics between the ion motion and
internal levels. Some of these dynamics for a single ion
have been discussed in Sec. 3.2. An interesting system
which can be simulated with these couplings is a
“phonon maser” which provides vibrational ampli-
fication by stimulated emission ofˇ radiation [309]. This
would be an analog of the micromaser with some
interesting differences such as the effects of recoil in the
pumping process.
The Hamiltonians of Eqs. (14), (37) and (48) for a
single ion can be consolidated into the general interac-
tion Hamiltonian
H
I
=
"V
S
e
+
e
[k ? xdt+
f
]
+h.c., (146)
where d =
v
v
0
, where we have used the rotating-wave
approximation, and where we have neglected phase
factors of the field. The parameter
e
is equal to 1 when
internal state transitions are involved and
e
= 0 when the
internal state is unchanged. In Eq. (146), k is the
wavevector of the field for single photon transitions or
k k
1
k
2
when stimulated-Raman transitions are
used. Similarly,
v
is the frequency of the applied field
for single-photon transitions and
v
v
L1
v
L2
when
stimulated-Raman transitions are used. In an interaction
picture of the ion’s motion, this Hamiltonian becomes
H
'
I
=
"V
F
S
e
+
e
i(dt–
f
)
P
j=x,y,z
exp(i[
h
j
(a
j
e
i
v
j
t
+a
j
e
i
v
j
t
)+h.c.
G
, (147)
where
h
x
k?xˆx
0
, a
x
is the lowering operator for the
x motion of frequency
v
x
, etc. Now, assume that
V
is
small enough, and that, in general,
v
x
,
v
y
,and
v
z
are incommensurate, so that we can excite only one
spectral component of the possible transitions induced
by this interaction. For a particular resonance condition
d
=–
,
x
v
x
,
y
v
y
,
z
v
z
, and in the Lamb-Dicke limit,
we find
H
'
I
"V
e
i
f
S
e
+
P
j=x,y,z
F
d
,
j
,u
,
j
u
(i
h
j
a
j
)
u
,
j
u
u
,
j
u!
+(1
d
,
f
,u
,
j
u
)
(i
h
j
a
j
)
u
,
j
u
u
,
j
u!
G
+h.c. , (148)
the two mode case where
e
=
,
z
= 0 is considered by
Drobny´ and Hladky´ [310], and in a different excitation
scheme in Ref. [110]. If the Lamb-Dicke limit is not
rigorously satisfied, we must consider higher order non-
linear corrections to this expression; specific examples
are discussed in Refs. [28] and [110]. These nonlinear
terms are the origin of the high-order corrections to the
Rabi frequencies (Eq. (18)). The case of carrier, and first
red and blue sidebands on internal state transitions (e.g.,
e
=1,
,
x
=
,
y
= 0,
,
z
=0,61) are used extensively for
quantum logic and generation of nonclassical motional
states and are discussed above. The case
e
=0,
,
x
=
,
y
= 0, u
,
z
u = 1 has been used to create coherent
[21] and Schro¨dinger cat [47] states of motion and is
discussed in Sec. 3.1. The case
e
=0,
,
x
=
,
y
= 0,
u
,
z
u = 2 has been used to create squeezed states; this is
discussed in Ref. [21]. A realization of the Hamiltonian
H
'
I
~ S
+
(a
x
)
2
+h.c. (
e
=1,
,
x
=
,
y
= 0,
,
z
= –2) has
been reported by Leibfried, et al. [132]. This is similar
to the case of two-photon excitation in cavity QED
analyzed by Buck and Sukumar [311] and Knight [312].
313
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
Interactions proportional to S
+
a
x
a
y
+h.c., S
+
a
x
a
y
+h.c.,
and a
y
a
z
+h.c. might be used to generate the maximally
entangled state of Eq. (60) without the need to address
individual ions [194]. An example of an interesting new
case would perhaps be the realization of three-phonon
downconversion (e.g.,
e
=0,
,
x
=3,
,
y
=–1,
,
z
= 0).
This case is analogous to three-photon downconversion
in quantum optics (see Refs. [110],[313], and references
therein). Here, it corresponds to driving a two-mode
resonance using stimulated-Raman transitions where
v
L1
v
L2
=
v
y
–3
v
x
. A suggestion to realize a
Hamiltonian ~a
x
2
a
y
+h.c. (
e
=0,
,
z
=0,
,
x
=2,
,
y
=–1)
is discussed by Agarwal and Banerji [314]. Clearly, a
large number of possibilities could, in principle, be
realized just for a single ion; moreover, the number of
possibilities increases dramatically if we consider all
modes of motion for multiple trapped ions. The only
limitation on how high u
,
j
u in Eq. (148) can be is that
V
be chosen sufficiently small that couplings to other
(unwanted) resonances are avoided. This will require
that decoherence be small enough to see the desired
dynamical behavior before coherence is lost.
Various forms of interactions which satisfy the
requirements of quantum nondemolition (QND) mea-
surements [20,25,315] of ion motion or quantum
feedback [316] can be extracted from Eq. (148). QND
experiments employing dispersive interactions are
considered by Retamal and Zagury [36]. These schemes
rely on measuring the ion’s internal state however,
which almost always involves recoil heating, thereby
destroying the state we wish to preserve. This could be
circumvented by coupling the ion’s motion to a cavity
field which then serves as the probe [317].
6.2.1 Mach-Zehnder Boson Interferometer With
Entangled States
Realization of the various Hamiltonians indicated in
Eq. (148) can lead to simulation of various devices of
practical interest. As an example, consider a Mach-
Zehnder interferometer which acts on two modes of
oscillation of a single trapped ion; to be specific, we will
consider the x and y modes of motion. The analogy with
a Mach-Zehnder interferometer for bosons is that
the two input modes to the boson interferometer are
replaced by the x and y modes of ion oscillation. The
(50/50) beamsplitters in the boson interferometer are
replaced by an operator B
6
= exp[6ip(a
x
a
y
+a
x
a
y
/4]
[193,318,319]. This operator can be realized by
applying the interaction in Eq. (148) with
e
=0, and
,
x
=–
,
y
= 1 for a time given by
Vh
x
h
y
t=p/4. A
differential phase shift between the two arms of the
interferometer can be simulated by shifting the relative
phases of the fields in Eq. (148) between successive of
applications B
6
. In a particle (e.g., boson) interfero-
meter, one typically measures the number of particles in
either one or both output modes. In a single ion experi-
ment, we have only one convenient observable, the inter-
nal state of the ion (either ul or ul. Nevertheless, we
can characterize the action of the phonon interferometer
by repeating the experiment many times and measuring
the density matrix of the output state [132].
It will be most interesting to characterize the action of
the interferometer for various nonclassical input states.
One example is the two-mode Fock state un
x
l
x
un
y
l
y
[195]. This state could be prepared by applying the
techniques described in Sec. 3.2 sequentially to the ion’s
x and y modes. This state is interesting because it has
been shown that one could approach the Heisenberg
uncertainty limit in a Mach-Zehnder interferometer by
measuring the distribution of bosons in the output
modes [195,196,197]. An alternative technique for
studying the action of a beamsplitter on the two-mode
Fock states has been suggested by Gou and Knight [23]
when
v
x
=
v
y
. Here, a beamsplitter could be simulated
by first preparing un
x
l
x
un
y
l
y
along two orthogonal axes
and then probing along two other axes (x' and y') which
are rotated (in the xy plane) with respect to the first. This
technique could also be used to analyze, for example, the
(u0l
x'
u2l
y'
+ u2l
x'
u0l
y'
)/Ï2 state from an initially pre-
pared u1l
x
u1l
y
state [23].
Another interesting state to consider for the phonon
interferometer is the (uNl
x
u0l
y
+(u0l
x
uNl
y
)/2
1/2
state
(which is the desired state after the first beam splitter).
This state has been shown to yield exactly the
Heisenberg uncertainty limit for an interferometer for
any value of N [194], if after the second beamsplitter,
we measure the number of particles N(x)inthex output
port. The result of this measurement is assigned the
value (–1)
N(x)
. This state could be prepared from the
ulu0l
x
u0l
y
state by the following two steps:
(1) Apply a p/2 pulse on the Nth blue sideband of mode
x [
e
=1,
,
x
=–N,
,
y
= 0 in Eq. (148)]; this creates the
state (ulu0l
x
+uluNl
x
)u0l
y
/2
1/2
.
(2) Apply a p pulse on the Nth blue sideband of mode
y [
e
=1,
,
x
=0
,
y
=–N in Eq. (148)]; this creates the
state ul(uNl
x
u0l
y
+u0l
x
uNl
y
)/2
1/2
.
After the second beamsplitter, we have a state which can
be written as
C
final
= ul
O
N
n
x
= 0
C
n
x
un
x
l
x
uN–n
x
)
y
. (149)
314
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
In principle, we would like to measure
ˆ
n
x
, record the
value N(x), and assign the value (–1)
N(x)
. Effectively,
this assignment can be accomplished if we can find an
interaction M which provides the mapping
M
C
final
= ul
O
N
n
x
even
C
n
x
e
i
f
(n
x
)
un
x
l
x
uN–n
x
)
y
+ ul
O
N
n
x
odd
C
n
x
e
i
f
(n
x
)
un
x
l
x
uN–n
x
)
y
. (150)
After this mapping, we need only measure the internal
state; if the ion is found in the ul state we assign the
value +1 to the measurement; if the ion is found in
the ulstate, we assign the value –1. The mapping
M can be achieved by applying radiation with k i xˆ
at the carrier frequency (
e
=1,
,
x
= 0) and insuring
V
n,n
t =2pm6n
x
p where m is an integer. From Eq. (18),
we have
V
n
x
,n
x
t
V
te
h
x
2
2
F
1–n
x
h
x
2
S
1+
h
x
2
4
n
x
h
x
2
4
DG
. (151)
Therefore, if we make
V
exp(–
h
x
2
/2)t=2mp and
h
x
2
(1+
h
x
2
/4=(2m
–1
) we achieve the desired mapping as
long as the contribution to the phase from the term
proportional to n
x
2
in this equation is small compared to
p. Therefore we require m>>N
2
/8 or, equivalently,
h
x
<<2/N.
One final example of a two-mode phonon interfero-
meter which directly yields Heisenberg 1/N phase sensi-
tivity is a “beamsplitter” which creates the state
(ulu0l
x
uNl
y
+uluNl
x
u0l
y
)/Ï2. This state can be cre-
ated by starting with an initial (ul+ul)uN/2l
x
uN/2l
y
/
Ï2 dual Fock state (Sec. 3.2). Next, we apply N/2
p-pulses alternating between the two interaction
Hamiltonians H
1
=
Vh
x
h
y
(S
+
a
x
a
y
+h.c.)and
H
2
=
Vh
x
h
y
(S
+
a
x
a
y
+h.c.) which can be realized in the
Lamb-Dicke regime. In this way, the state of motion of
the ion is stepped through
ulu
N
2
l
x
u
N
2
l
y
+ulu
N
2
l
x
u
N
2
l
y
H
1
ulu
N
2
+1l
x
u
N
2
–1l
y
+ulu
N
2
–1l
x
u
N
2
+1l
y
H
2
ulu
N
2
+2l
x
u
N
2
–2l
y
+ulu
N
2
–2l
x
u
N
2
+2l
y
H
1
.
. (152)
.
uluNl
x
u0l
y
+ulu0l
x
uNl
y
where, in this example, we assume N/2 is even. The
interactions H
1
and H
2
follow from Eq. (148) with e=1,
,
x
=–1,
,
y
= 1,
,
z
= 0 and e=1,
,
x
=1,
,
y
= –1,
,
z
=0,
respectively. The kth pulse has Rabi frequency
Vh
x
h
y
[(N/2+k)(N/2–k+1)]
1/2
in the Lamb-Dicke re-
gime. After a relative phase is accumulated in the two
“paths” of the interferometer (simulated by adjusting
the phase of the laser pulses as discussed above), we
reverse the above steps and apply a final “p/2” pulse on
the carrier. Upon measuring the probability of occu-
pation in state ul or ul, the interference fringes exhibit
1/N phase sensitivity.
If the Lamb-Dicke criterion is not satisfied, the two
components of the wavefunction superposition may
experience different Rabi frequencies during each pulse,
leading to undesired evolution. The exact Rabi frequen-
cies of the two components of the wavefunction upon
application of the kth pulse (interaction H
1
or H
2
) follow
from Eq. (18):
V
A
=
Vh
x
h
y
e
–1/2(
h
x
2
+
h
y
2
)
L
N/2+k–1
1
(
h
x
2
) L
N/2–k
1
(
h
y
2
)
Ï(N/2+k)(N/2–k+1)
,
(153)
V
B
=
Vh
x
h
y
e
–1/2(
h
x
2
+
h
y
2
)
L
N/2+k–1
1
(
h
y
2
) L
N/2–k
1
(
h
x
2
)
Ï(N/2+k)(N/2–k+1)
,
where
V
A
is the Rabi frequency between states
uN/2+k –1l
x
|N/2k+1l
y
and uN/2+kl
x
|N/2kl
y
and
V
B
is the Rabi frequency between states uN/2–k+1l
x
|N/
2+k –1l
y
and uN/2–kl
x
|N/2+kl
y
. The only differences
between
V
A
and
V
B
are the arguments of the associated
Laguerre polynomials. Thus, as long as
h
x
=
h
y
, the
system will evolve as in Eq. (152), even when the Lamb-
Dicke criterion is not satisfied.
6.2.2 Squeezed-Spin States
As an example which demonstrates the advantages of
using entangled states in spectroscopy, we discuss the
following simple experiment which can be carried out
on a single ion. Let us suppose we are interested in
measuring, with maximum signal-to-noise ratio, the
Zeeman frequency between states of a J = 1 manifold in
an atom. To be specific, consider that we are interested
in measuring the Zeeman frequency of the DM
F
= 61
transitions in the
2
S
1/2
(F=1) ground-state hyperfine
multiplet of a
9
Be
+
ion (Figs. 5 and 11). This manifold
is composed of the three levels uF,M
F
l [ {u1,0,l,
u1,61l}. We will assume that the applied field is small
enough that the frequency of the u1,–1l u1,0l transi-
tion is equal to the frequency of the u1,0l u1,1l transi-
tion. A straightforward way to measure the Zeeman
frequency would be to prepare the atom in the u1,1l
315
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
(or u1,–1l state, drive the Zeeman transition (using the
Ramsey method), and then measure J
z
. We will assume
this detection can be accomplished with 100 % detec-
tion efficiency, so the noise in the measurement is
limited by the quantum statistics in the measurement
process [100]. After many measurements, requiring a
total averaging time
t
, a certain precision in the Zeeman
frequency would be obtained. However an alternative
measurement strategy, described below, would require
an averaging time
t
/2 to reach the same measurement
precision.
The basic idea is outlined in Ref. [9]; for J =1,itis
equivalent to the more general technique described by
[194]. We prepare the atom in the state
c
(0) = u1,0l and
then apply the Ramsey fields. Subsequently, we measure
the probability of the ion to remain in the u1,0l state.
(This measurement scheme has been used in the exper-
iments of Abdullah et al. [320]; the noise in these exper-
iments, however, was not limited by projection noise.)
This measurement is equivalent to measuring the opera-
tor
II
J
2
z
, where
II
is the identity operator, and therefore
equivalent to measuring the operator J
2
z
, as discussed by
Wineland, et al. [9], or the operator P
i
s
zi
, as discussed
by Bollinger, et al. [194]. After application of the first
Ramsey pulse to the u1,0l state, the ion is in the state
(u1,1> + u1,–1l)/Ï2; this is equivalent to the maximally
entangled state of Eq. (60) which could be formed by
two spin-1/2 particles. As described in Sec. 3.3, we
would expect the time required to reach a certain
measurement precision to be reduced by a factor of
2(L = 2) over the case of uncorrelated particles, repre-
sented by starting the ion in the state
c
(0) = u1,1l or
u1,–1l state and finally measuring J
z
.
We could carry out this experiment in the following
way. We first optically pump the ion to the u2,2l hyper-
fine state. We then apply two successive p pulses which
carry out the transformations u2,2l u2,1l u1,0l with
stimulated-Raman transitions. We then perform Ramsey
spectroscopy on the u1,0l state. We can measure the
probability of subsequently finding the atom in the u1,0l
state by first reversing the order of the two p pulses and
then measuring the probability of finding the ion in the
u2,2l state.
6.3 Mass Spectroscopy and NMR at the Single
Quantum Level
Quantum logic operations may be useful in precision
measurements other than spectroscopy (Sec. 3.4). For
example, Ref. [48] discusses a method to measure
cyclotron resonance frequencies of single ions at the
quantum level; this technique essentially employs quan-
tum logic to distinguish between motional quantum
states. This capability would provide mass spectroscopy
at such low energies that anharmonic shifts (electric
potential and relativistic) would be very small. (At
present, however, high-resolution mass spectrometry
experiments are limited by other effects such as mag-
netic field drifts [77,78,79].) In mass spectroscopy, the
basic problem is to measure a “test” ion’s cyclotron
frequency and compare it to the cyclotron frequency of
a “measurement” ion (in the same magnetic field) which
serves as the transfer standard.
One idea [48] is to store two ions in separate Penning
traps which are stacked along their symmetry axis and
share a common endcap and a common, axial magnetic
field. The axial frequencies of the ions are adjusted to be
the same, in which case the axial harmonic oscillators
are coupled through the charge in the common endcap.
The technique might work as follows: The axial motion
of the “measurement” ion (for example,
9
Be
+
)isfirst
laser cooled to the n = 0 level. When the ions’ axial
modes are resonantly coupled, this cooling can be trans-
ferred to the test ion’s axial motion; if the coupling is left
on for a certain amount of time, the energies in the ions’
axial oscillations are exchanged. Subsequently, the
9
Be
+
ion’s axial oscillation is recooled so that both ions are
cooled to the zero-point energy. This axial cooling is
then transferred to the test ion’s cyclotron mode by
parametric coupling, after which the axial motion of
the measurement ion must be recooled to the zero-point
energy. An external field is then applied to weakly
excite the test ion’s cyclotron motion. When the ampli-
tude of this field is adjusted appropriately and the
resonance condition is met, the n =0ton = 1 transition
in the test ion is driven with high probability. The steps
above are then reversed so that if the test ion’s cyclotron
motion was excited to the n = 1 level, the
9
Be
+
ion’s axial
motion is now in the n = 1 level. When the resonance
condition for the test ion’s cyclotron frequency is not
met, the
9
Be
+
ion’s motion remains in the n = 0 level.
Discrimination between the n = 0 and n = 1 axial level is
then performed using quantum logic operations applied
to the
9
Be
+
ion as discussed in Sec. 3. For example, after
the steps above, an axial red sideband p pulse will excite
a ul ul transition in the
9
Be
+
ion conditioned on
whether or not the
9
Be
+
ion’s axial mode was in the
n =0orn = 1 state. The test ion’s cyclotron frequency
can be referenced to a spin flip frequency in
9
Be
+
(either
electron or nuclear) which then acts as a transfer
standard. Finally, by performing a cyclotron resonance
measurement on a second test ion in the same fashion
we can find the ratio of the two test ions’ cyclotron
frequencies and therefore derive their mass ratio.
Ref. [48] also suggested that these ideas could be
applied to measure magnetic moments of test ions. In
this section, we describe a variation on the method dis-
cussed there. The basic idea is to perform NMR on an
––
––
316
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
unknown magnetic moment (the test ion) at the single
spin level. Using quantum logic operations, the spin flip
is detected in a measurement ion to which the test ion is
coupled. We will describe the ideas in the context of a
specific example; but the techniques are easily general-
ized to measuring the magnetic moments of other ions.
Suppose we want to measure the ratio of proton and
antiproton magnetic moment. Such a measurement is of
high current interest and can provide a second test of
CPT on baryons in addition to the precise mass com-
parisons already performed [79]. Precision measure-
ments of the magnetic moment ratios coupled with the
mass ratios should test much of the same physics in the
context of CPT as precision spectroscopy of antihydro-
gen [321]. A measurement scheme for the proton/
antiproton moment ratio has been suggested previously
by Quint and Gabrielse [322]. This method would
employ the same basic ideas as in the electron g-2
experiments of Dehmelt and coworkers [108], where the
proton’s (or antiproton’s) magnetic moment energy is
transferred parametrically into its cyclotron energy
(which is then detected) by use of applied inhomoge-
neous (oscillating) magnetic fields. The apparent draw-
back to this scheme is that it is slow because of the
weakness of the parametric coupling and the difficulty
of detecting small changes in the cyclotron energy. The
method suggested here incorporates the same kind of
parametric coupling but is potentially more efficient
because it is sensitive to transfer at the single quantum
level through the use of quantum logic techniques.
The basic idea of the method we propose here is to
first store a proton and a
9
Be
+
ion in trapping potentials
which are in close proximity and in a common mag-
netic field (Fig. 12). The proton spin flip frequency is
then compared to the
9
Be
+
spin flip frequency (elec-
tron or nuclear), which acts as a transfer standard, effec-
tively calibrating the (common) magnetic field. We then
perform the same kind of measurement on a simulta-
neously trapped antiproton and
9
Be
+
ion. By combining
the two measurements, we determine the proton/
antiproton magnetic moment ratio.
First consider the measurement of the proton to
9
Be
+
spin flip frequency ratio. We simultaneously store a
single proton and
9
Be
+
ion in a double-well potential as
indicated in Fig. 12. We assume that the ions are con-
fined in a direction perpendicular to the z axisbya
linear rf trap combined with a superimposed static
magnetic field B = B
0
zˆ. (One particular geometry for a
Paul trap in a strong magnetic fields is described by
Bate, et al. [323].) The quadrupole electrodes are
segmented to provide a static double-well potential
along the z direction. We assume the electrode segments
are capacitively coupled together so the rf electrode
potentials are independent of z. A vertical wire (shown
in cross section in the figure) is at one end of the trap;
oscillating currents in this wire generate an oscillating
magnetic field which can be used to drive the spin flip
transitions of the proton and nuclear spin flip transition
in the
9
Be
+
ion (the electron spin flip transition in
9
Be
+
could be driven with injected microwave radiation).
Oscillating currents in this wire will also provide a
parametric coupling between the proton spin and axial
motion (below).
The experiment could proceed as follows: We first
tune the static potentials so that the proton and
9
Be
+
axial frequencies are the same. By performing Doppler
laser cooling on the
9
Be
+
ion’s axial and radial modes,
the proton’s axial motion is cooled. The proton’s radial
modes will also likely be cooled through the Coulomb
coupling. If not, we must apply an inhomogeneous
rf field which parametrically couples the proton’s radial
and axial frequencies. We now uncouple the proton
and
9
Be
+
ion by adjusting their axial frequencies to
be different. We then cool the
9
Be
+
axial motion to the
ground state (Sec. 3.1) and follow this by switching
the axial modes back into resonance. The Coulomb
interaction between the ions gives rise to a coupling
Fig. 12. Schematic diagram of a trap for simultaneous storage of two
ions with different charge to mass ratios, q
1
/m
1
and q
2
/m
2
. We assume
the trap is generically equivalent to the linear trap of Fig. 1. The rods
of the trap in Fig. 1 are replaced by segmented electrodes in which
adjacent segments are at the same rf potential but where the segments
support different static potentials in order to make a double well
potential along z. Therefore in this figure, the view is from above,
where we see two (of four) segmented electrodes which have replaced
two of the rods in the trap of Fig. 1. On the left is a cross section of
a wire through which an rf oscillating current is sent; this wire gener-
ates an rf magnetic field at the site of the ions as explained in the text.
317
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
q
1
q
2
/(4p
e
0
uZ
2
–Z
1
u), where Z
1
and Z
2
are the respective
axial positions of the ions and q
1
and q
2
are, here, the
charges of the proton and
9
Be
+
ion. We have neglected
the effects of induced charges in the trap electrodes, this
will not be a large effect if the distance between ions is
on the order of or less than the distance of each ion to
the trap electrodes. For small amplitudes of oscillation
we can write Z = Z
i0
+z
i
, i [ {1,2} where Z
i0
are the
equilibrium positions of each ion. In the limit of weak
coupling and to lowest order in z
i
, the interaction
between ions is given by
H
q
1
,q
2
=
q
1
q
2
z
1
z
2
2p
e
0
d
3
, (154)
where d uZ
10
–Z
20
u. This gives rise to a frequency split-
ting between axial normal modes of [48]
d
v
1,2
= p/t
exch
=
q
1
q
2
2p
e
0
d
3
v
z
Ïm
1
m
2
, (155)
where m
i
are the ion masses and t
exch
is the time to
exchange the energy of the two ions’ axial energies after
they are coupled together. At the time t
exch
, the ions are
again decoupled, leaving the proton in its axial ground
state. The
9
Be
+
ion is subsequently recooled to its axial
ground state.
Assume the proton is initially in the lower energy
state |l. We now apply an oscillating magnetic field at
a frequency
v
m
which is near the proton spin flip
frequency
v
0
= g
p
m
B
B
0
/
"
where g
p
is the proton gyro-
magnetic ratio expressed in terms of the Bohr magneton
(g
p
1.521310
–3
). The amplitude and duration of this
field is adjusted to make a p pulse if the resonance
condition
v
m
=
v
0
is satisfied; that is, the proton spin
undergoes the transition ul ul. We now apply an
inhomogeneous magnetic field at a frequency
v
a
by
driving a current Icos
v
a
t through the vertical wire,
shown on the left in Fig. 12. If we neglect the shielding
effects of the trap electrodes, this provides a field at the
site of ion 1 equal to B(t)=xˆ
m
0
Icos
v
a
t/(2p
r
)anda
field gradient B
x
/ z =–Icos
v
a
t/(2p
r
2
) where
r
is the
distance from the wire to the proton. This oscillating
field gives rise to a coupling which has the form of
Eq. (28). If the resonance condition
v
a
=
v
0
v
z
is satis-
fied, this coupling takes the form of Eq. (27) with
V
1
= g
p
m
B
m
0
Iz
0
/(4p
"r
2
) where z
0
is the zero-point
motion for the proton (as in Sec. 2) . We assume that the
duration t
coupling
of this inhomogeneous field is adjusted
to give a complete transfer ulu0l ulu1l(
V
1
t
coupling
=
p/2). We now couple the axial modes for a time t
exch
so
that if the proton was in the un=1l state after the last
step, the
9
Be
+
ion is now in the un =1l state. Finally, the
un =1l motional state is detected on the
9
Be
+
ion using
the methods as discussed above and in Sec. 3. If
v
m
is
nonresonant, the modes remain in the un =0l state.
If the proton is initially in the ul state and
v
m
is
resonant, then the field at frequency
v
a
=
v
0
v
z
has no
effect, and the ion remains in the un =0l motional state
and gives a false “no signal.” However, when
v
m
is
reapplied, the un =1l state is produced giving a signal.
After reinitializing (that is, preparing both ions in their
axial ground states), signals are always produced if
v
m
remains resonant. If
v
m
is nonresonant and the proton
initial state is ul, no signals are ever produced. However
if
v
m
is nonresonant and the proton initial state is ul,we
produce a false signal. After reinitializing and repeating
the experiment, we produce no signals. This is true on
subsequent tries as long as
v
m
remains nonresonant.
Therefore to reduce the effects of false signals, we
should repeat each try several times for each value of
v
m
and discard the first measurement. Interlaced with
measurements of the proton spin flip frequency, we
measure the
9
Be
+
spin flip frequency by driving with
resonant rf radiation and using the techniques outlined
in Sec. 3 for detection. This allows us to monitor and
correct for magnetic field drifts on a fairly short time
scale.
The antiproton/
9
Be
+
comparison is accomplished
similarly except for the important difference that axial
potentials must be trapping for one species and nontrap-
ping for the other at a given location. Apparently, the
biggest liability in the scheme presented here is the
same as that of the proposal of Quint and Gabrielse
[322], namely, the weakness of the parametric coupling
between proton magnetic moment and axial motion.
For
v
z
/2p =1 MHz, we find z
0
(proton)=71 nm. With
I=1 A, and
r
= 1 mm, we find
V
1
/2p 0.15 Hz. For
these same conditions and d = 0.5 mm, we have
t
exch
27 ms. At a field of 5 T,
v
0
/2p 106 MHz. At
very high resolutions we want resonance linewidths of
less than 1 Hz which, in turn, requires long resonance
times for the proton spin flip. Therefore the long times
required for t
coupling
need not be a serious liability.
We have not considered details of proton, antiproton,
or
9
Be
+
transfer into the trap, however this might be
accomplished by adapting a scheme similar to that
described in Sec. 4.1. The trapping arrangement we
show in Fig. 12 is essentially the same as the coupled
trap idea of Ref. [48]. By removing the common endcap
in the coupled-trap scheme, we arrive at the situation
depicted in Fig. 12. In either case, we can show that the
coupling between ions is approximately given by Eq.
(154) where d is the overall distance between ions 1 and
2. The potential advantage of the scheme described
here is that d can probably be made smaller than
in a coupled trap, thereby reducing t
exch
.Wehave
assumed the use of
9
Be
+
ions, but many other ions
318
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
would work. A potential advantage of
9
Be
+
ions is that
the axial potential wells required to make
v
1
=
v
2
are
not as different as for other choices of ion 2. In a similar
spirit, we could measure the proton (or antiproton)
magnetic moment in terms of other atomic parameters.
For example, the proton spin flip frequency vs
9
Be
+
spin
flip frequency could be combined with a separate
measurement of the
9
Be
+
spin flip frequency compared
to its cyclotron frequency to yield a measurement of the
proton spin flip frequency to
9
Be
+
cyclotron frequency
in the same magnetic field. Coupled with an accurate
value of the electron to
9
Be
+
mass ratio, these measure-
ments yield an accurate value of g
p
. An important
systematic effect to consider in these measurements is
magnetic field inhomogeneity. Field homogeneity could
be checked to high accuracy by moving the
9
Be
+
ion to
various locations in the trap, thereby mapping the
magnetic field. Equally important in the proton/antipro-
ton comparison is to insure that the proton and anti-
proton are in the same location. It appears this could be
accomplished by insuring the
9
Be
+
is always at the same
location (using optical means) and adjusting the trap
potentials to always yield the same values of
v
z
and
d
v
1,2
. By this method, it appears that accuracy signifi-
cantly better that 1 part in 10
9
could be achieved.
6.4 Quantum State Manipulation of Mesoscopic
Mechanical Resonators
Much of what we have discussed concerns manipula-
tion of the mechanical oscillation of atoms or atomic
ions. It is perhaps interesting to speculate on the possi-
bility of applying similar techniques to the manipulation
of macroscopic or mesoscopic mechanical resonators in
the quantum regime [119,120]. If studies can be per-
formed at the quantum level, new sensors at the single
phonon level could be built [324]. One approach is to
make the mechanical resonators small enough and
the temperature low enough that
"v
> k
B
T [325]; this
approach may obviate the idea discussed here. The idea
we examine here is an extension of the idea of coupling
the oscillatory motion of two ions together; here we
consider coupling the motion of an ion (or the COM
mode of a collection of ions) to a single mode of
mesoscopic mechanical resonator. Although current
technology appears to prohibit performing such experi-
ments at the quantum level, the development of
mesoscopic resonators is rapid and such experiments
may be possible in the future. Similar considerations
regarding coupling of ions to piezoelectric resonators
were discussed in Ref. [48].
To be specific, we consider the situation sketched in
Fig. 13. The mechanical resonator is assumed to be
a silicon beam resonator, fixed at both ends; we will
assume the conditions realized in the experiments of
Cleland and Roukes [325]. The beam resonator has a
length D, thickness in the z direction equal to Dz and a
thickness in the x direction (out of the plane of Fig. 13)
equal to Dx. If we take the conditions of Fig. 3 of
Cleland and Roukes [325] as a guide, we have
D = 7.7 mm, Dz = 0.8 mm, and Dx = 0.33 mm. We
assume that the beam is metalized near its center and
can support a charge q
1
. For simplicity, we will assume
the metalization is confined to a spherical shell of radius
R
1
, but the exact geometry is not so important. Neglect-
ing the dielectric effects of the Si beam (T = 4 K), the
capacitance of the metallized sphere is approximately
equal to C
b
=4p
e
0
R
1
. The sphere can therefore support
a charge q
1
= C
b
V
b
where V
b
is the potential on the
sphere. An atomic ion of charge q
2
and mass m
2
is
trapped by a combination of electrodes at a distance d
from the cantilever. One of the electrodes is the
charged, metallized sphere; the other electrodes are
indicated schematically as A, B, and C. To estimate the
mass m
2
of the mechanical resonator we assume it has
a mass equal to half of the beam’s mass and that this
mass is concentrated in the metallized sphere. For the
conditions assumed here, and
r
(Si) = 2.33 g? cm
–3
,we
find m
1
2.4310
–12
g. We arrange the trapping poten-
tials so that the z-oscillation frequency of the ion is
equal to the beam oscillation frequency; in this case, we
realize two coupled oscillators as described in the
previous section. If we take d =5mm, V
b
= 1000 V,
v¯
/2p = 70 MHz, q
2
= q(proton), m
2
=9u (
9
Be
+
), we
find from Eq. (155) that t
exch
6.4 ms. To reliably work
in the quantum regime of the mechanical oscillator, t
exch
Fig. 13. Schematic diagram of an ion coupled to mechanical
resonator. The bridge-type cantilever is essentially the same as that
reported by Cleland and Roukes [325], but is assumed to support a
metallized sphere supporting charge q
1
.Anionq
2
is confined by the
potentials on the metallized sphere and additional electrodes shown
schematically as A, B, and C. The Coulomb coupling between the
charges provides the coupling between the two harmonic oscillators.
319
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
must be smaller than t
*
, the time for the mechanical
oscillator to make a transition from its un =0l ground
state to the un =1l state. If we assume the conditions of
Cleland and Roukes [325] where T =4K, and
Q =2310
4
, Eq. (64) gives t
*
0.0382 ms. This is
clearly too short to sympathetically cool the beam
resonator’s mode to the ground state, although some
cooling could be achieved as outlined by in Ref. [48]. In
the future, it might be possible to make the ratio t
*
/t
exch
larger than 1, perhaps through higher Qs and lower
ambient temperatures [or perhaps if more exotic ion
species with Z
2
>> 1 become available (Sec. 5.2)], it
may be possible to manipulate the quantum motion of
mesoscopic mechanical resonators by these techniques.
7. Summary/Conclusions
We have attempted to identify some of the important
practical effects which must be taken into account in
order to create arbitrary, entangled quantum states of
trapped ions. We have taken a “passive stabilization”
approach in that we try to anticipate and suggest ways to
guard against the physical effects causing decoherence.
Ultimately, complicated manipulations, such as lengthy
quantum computations, are expected to benefit from
and/or require some form of active error correction.
Indeed, some of the near-term future experiments will
probably demonstrate some of these schemes. In
Sec. IV, we have listed some of the potential sources of
decoherence in trapped ion experiments; here we
speculate on what appear to be the most important of
these.
Motional decoherence is discussed in Sec. 4.1.
In the NIST single
9
Be
+
ion experiments [17,21,45,47,
131,132,211], heating appears to be the most important
source of decoherence, primarily because, at the present
time, its source is still unidentified. Various possibilities
were discussed and, although the heating may be caused
by some fundamental effect, we speculate that it is
probably caused by some, as yet undetected, source of
added electric field noise. (Because the current experi-
ments employ electronic filtering at the motional
frequencies, direct observation of fluctuating potentials
on the electrodes has been precluded.) Future experi-
ments will be able to resolve this. Moreover, once this
source of noise is understood, the ion becomes an
extremely sensitive detector of fluctuating potentials
appearing on the electrode surfaces. In any case, it will
probably be desirable to eventually operate the ion
trap at cryogenic temperatures in order to, for example,
reduce the effects of ion loss due to background
gas collisions. A cryogenic environment will have the
added benefit of reducing sources of electronic noise
and associated heating.
Various sources of internal-state decoherence are
discussed in Sec. 4.2. In the current experiments,
decoherence is dominated by fluctuations in magnetic
fields acting on qubit transitions which are strongly field
dependent. It appears that this problem can be highly
suppressed by the use of magnetic shielding and, eventu-
ally, use of qubit transitions which are insensitive
to magnetic field to first order. We expect that internal
state decoherence will be negligible compared to
motional decoherence.
Decoherence induced during application of the logic
pulses (Sec. 4.4) may be the most troublesome. Many of
the sources of this type of decoherence are primarily
technical, for example, caused by intensity fluctuations
in the laser pulses which induce transitions, or
stray light impinging on ions not directly addressed.
The more fundamental causes of decoherence are
(1) coupling to internal or motional states other than the
intended ones (spectator level problem), (2) coupling to
unintended motional modes (cross-mode coupling), and
(3) fluctuations in the Rabi rates due to excitation of the
3L-1 unused motional modes (Debye-Waller factors).
The first two of these effects appear to be a question of
speed. Coupling to unwanted spectator modes and
cross-mode coupling can always be avoided by making
the operations slow enough that the extraneous
couplings are suppressed by spectral selection. This has
the negative effect of allowing more time for decoher-
ence and increasing the required time for a given
computation. Fluctuations caused by fluctuating Debye-
Waller factors are, in principle, reduced as the number
of ions increases because of the averaging effects of
many modes [Eq. (128)]. However, it is also likely that
excitation of these modes is harder to avoid as the
number of ions (and unwanted spectator modes)
increases. It will therefore be desirable to laser-cool all
modes to the zero-point state.
As the number of trapped ions increases, it will be-
come increasingly difficult to avoid these three types of
effects. Therefore, some sort of multiplexing scheme
will be necessary when large numbers of qubits are
involved. In Sec. 5.1, we have presented one possible
solution where the ions are broken up into smaller
numbers of independent groups or registers. The ions
are then connected by moving ions around between
registers. It may also be advantageous to multiplex
quantum information within multiple internal states of
ions; this is briefly described in Sec. 5.2.
Stimulated-Raman transitions between long-lived
qubit states (such as ground-state hyperfine levels) ap-
pear to offer significant advantages over single-photon
optical transitions. Single-photon transitions require
high laser frequency stability, whereas stimulated-
Raman transitions require only high relative frequency
320
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
stability between the Raman beams, which is techni-
cally easier to accomplish. Stimulated-Raman transi-
tions also provide the ability to select the magnitude and
direction of the effective k vector (k
eff
) by choosing
different directions for the beams. This has the advan-
tage of, for example, allowing k
eff
to be parallel to the
axis of ions in a linear trap, thereby suppressing
coupling to radial modes, while still allowing the Lamb-
Dicke parameter to be controlled by adjusting the angle
of the beams. At the same time, spatial selection of ions
along the axis of the trap can be good since each Raman
beam can be at a relatively steep angle relative to the
trap axis. With single photon transitions, selection of
ions and modes by appropriately choosing the direction
of k can only be obtained for the radial modes in a linear
trap. Stimulated-Raman transitions have the potential
disadvantage of inducing significant ac-Stark shifts
(Sec. 4.4.3). However, for anticipated operating condi-
tions, the effects of Stark shifts can be suppressed
relative to the effects of laser amplitude fluctuations,
which affect both single-photon and stimulated-Raman
transitions.
Decoherence in the ion trap system can probably be
substantially reduced over what has been obtained in
experiments so far. How far this reduction can be carried
is an issue which must be resolved experimentally.
Decohering effects may eventually be controlled to such
a level that fault-tolerant error correction schemes might
be employed to achieve computations of arbitrary
length. This may only require a single operation fidelity
of 10
–5
to 10
–6
[185,326]. If this condition is met, speed
will become an important issue because of the poten-
tially large amount of overhead (increase in number of
required qubits and operations in fault tolerant
schemes). As discussed in Sec. 4.4.6, the Rabi rate for
any operation is limited to approximately the motional
mode frequency. In principle, mode frequencies can be
substantially increased beyond what is currently
achieved (10 MHz). An obvious direction to pursue is
to make smaller traps with higher trapping potentials;
however, this aggravates the problem of addressability,
and will increase the coupling of the ions to the elec-
trodes thereby increasing decoherence. The optimum
conditions must, again, be resolved experimentally.
8. Appendix A. Entangled States and
Atomic Clocks
We compare the use of entangled vs nonentangled
states in an atomic clock under conditions and con-
straints different than those considered in Sec. 3.4. We
assume that the resources available are a given number
of atoms L and a total observation time which is longer
than any other time scale in the problem. We assume
that decoherence of the internal states is negligible
during the Ramsey free precession time T
R
.Tomakean
atomic clock, we want to steer, or “lock,” a reference
oscillator ( “local oscillator”) to the atoms’ resonance.
Typically, we can find a local oscillator whose rms
frequency fluctuations D
v
LO
(
t
) over short averaging
times
t
are smaller than the intrinsic fluctuations given
by projection noise. However, for long times, the
frequency fluctuations of the local oscillator are worse
than those given by projection noise. (If this were not the
case, the local oscillator would be a better clock, obviat-
ing the need for the atoms.) By measuring the atomic
populations after each application of the Ramsey radia-
tion [3,9,100], we can generate an error signal which
steers the local oscillator to the center of the atomic
resonance with a servo time constant
t
s
. The servo can
make a correction after a few independent measure-
ments on the atoms; for simplicity, we assume
t
s
~ T
R
(Eq. (A4) below). To make the error signal as large as
possible, we want the atomic linewidth D
v
a
as small as
possible. The atomic linewidth can be expressed as
D
v
a
= p/(L
(2
e
–1)
T
R
), where
e
= 1/2 for nonentangled
states and
e
= 1 for entangled states [3,9,100]. There-
fore, we want T
R
as large as possible. However, if we
make T
R
too large, the local oscillator fluctuations
D
v
LO
(T
R
) will be larger than D
v
a
, thereby giving no
useful signal. This is the constraint which tells us
whether entangled or nonentangled states are more
useful.
To analyze this problem in detail, we make the follow-
ing assumptions: (1) We assume the (free-running
or unlocked) local oscillator has fluctuations over
averaging time
t
equal to
D
v
LO
(
t
)=C
t
n
. (A1)
This is assumed to hold over the range of values of
t
which include T
R
and
t
s
. We will consider only values of
n > 1/2, since, for n < –1/2, the local oscillator will
better than the atoms for
t
sufficiently long. (2) We
assume the intrinsic atomic clock stability, limited by
projection noise, is given by (Sec. 3.4)
D
v
meas
(
t
)=
L
e
ÏT
R
t
, (A2)
(3) For a given servo time constant
t
s
, we assume
D
v
LO
(
t
s
)=K
1
D
v
meas
(
t
s
) (A3)
Normally, in the case where D
v
LO
< D
v
meas
for short
times
t
, we would think of adjusting
t
s
so that K
1
< 1;
that is, the local oscillator is locked to the atomic reso-
321
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
resonance at or before the time
t
when the local oscil-
lator fluctuations become worse than the projection
noise. However, when K
1
> 1, the locked local oscillator
stability can eventually reach the stability given by
projection noise (approaching it as 1/
t
) given adequate
servo gain [327-329]. Therefore we will allow K
1
to be
larger than 1.
Along with these assumptions, we impose two con-
straints: (1) The servo time constant
t
s
must be longer
than the sampling time T
R
. (We will assume T
R
is much
larger than the dead time, that is, the time for optical
pumping, detection, etc.) This is expressed by the condi-
tion
t
s
= K
2
T
R
, K
2
> 1. (A4)
For entangled states, each measurement gives one of two
possible values (Sec. 3.4). Therefore, K
2
must be larger
for entangled states than nonentangled states. (2) For
t
T
R
, the atomic linewidth must be greater than the
local oscillator fluctuations or else the error signal used
to correct the local oscillator frequency is ambiguous.
This is expressed by the condition
D
v
a
(T
R
)=p/(L
(2
e
–1)
T
R
= K
3
D
v
LO
(T
R
), K
3
>1.
(A5)
Eq. (A5) gives the sampling time T
R
as a function of K
3
T
R
=
S
p
CK
3
L
2
e
–1
D
1
n+1
. (A6)
Plugging this into Eq. (A2), we find that the stability of
the locked local oscillator at long times
t
is
D
v
meas
=
F
CK
3
p
G
1
2(n+1
L
(n
e
+1/2)
n+1
1
Ï
t
. (A7)
This is the main result of Appendix A. From this expres-
sion, we see that D
v
meas
is smaller for entangled states
when n > 0; however, the gain is not as significant as
when we assumed T
R
to be fixed. We also see that
D
v
meas
is smaller for nonentangled states when n <0.
These results are due to the constraint expressed in
Eq. (A5). From the above, we also find
K
1
=
pK
2
n+1/2
K
3
L
1
e
~ K
2
n+1/2
L
1
e
. (A8)
For nonentangled states on large numbers of atoms, K
1
and T
R
are both much larger than for entangled states,
and a very long time may be required to achieve the
intrinsic atomic clock stability given by Eqs. (A2) and
(A7). therefore, as a practical issue, we may wish to
constrain K
1
to be equal to 1. In that case we find, from
Eqs. (A3) and (A5) two values of R
R
. We must take the
smaller of these, which results from the value derived
from Eq. (A3) (with K
1
= 1). Plugging this value of T
R
into Eq. (A2), we find the stability of the locked oscilla-
tor to be
D
v
meas
=
F
CK
2
n+1/2
G
1
2(n+1)
L
e
S
(2n+1)
2n+2
D
1
Ï
t
. (A9)
For large L, entangled states will give a smaller value of
D
v
meas
for all values of n > –1/2; however this result is
not fundamental and simply comes from the constraint
that K
1
is equal to 1.
9. Appendix B. Master Equation for the
Density Matrix of a Radiatively
Damped Harmonic Oscillator
Equation (62) is the master equation for the density
matrix
r
of a single harmonic oscillator, including
radiative damping terms. This equation is equivalent to
the following system of coupled first-order differential
equations for the diagonal and off-diagonal number-
state matrix elements of
r
:
knu
r
?
unl
r
?
nn
=
g
(n
_
+1)(n+1)
r
n+1 n+1
g
(2nn
_
+n+n
_
)
r
nn
+
g
n
_
n
r
n–1 n–1
kmu
r
?
unl
r
?
mn
=
g
(n
_
+1)Ï(m+1)(n+1)
r
m+1 n+1
+
g
n
_
Ïmn
r
m–1 n–1
1
2
g
[2n
_
(m+n+1)+(m+n)]
r
mn
(B1)
Some special cases of these equations were given in Eq.
(63), where they were used to estimate the time for an
ion to make a transition from the state |0l. Cohen-
Tannoudji [330] treated a similar system, in which the
incoherent excitation was absent (n
_
= 0), but a
monochromatic perturbation was present. The time
derivative of the thermal average of the number of
vibrational quanta is
322
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
dkn
ˆ
l
dt
O
`
n=0
n
r
?
nn
=
g
(n
_
+1)
O
`
n=0
n(n+1)
r
n+1 n+1
g
n
_
O
`
n=0
n
r
nn
g
(2n
_
+1)
O
`
n=0
n
2
r
nn
+
g
n
_
O
`
n=1
n
2
r
n–1 n–1
(B2)
The sums can be simplified to
O
`
n=0
n(n+1)
r
n+1 n+1
=
O
`
n=0
(n+1–1)(n+1)
r
n+1 n+1
=
O
`
n=0
(n+1)
2
r
n+1 n+1
O
`
n=0
(n+1)
r
n+1 n+1
=
O
`
n=0
n
2
r
nn
O
`
n=0
n
r
nn
(B3)
and
O
`
n=1
n
2
r
n–1 n–1
=
O
`
n=1
[(n –1)
2
+2n–1]
r
n–1 n–1
=
O
`
n=1
[(n –1)
2
+2(n–1)+1]
r
n–1 n–1
=
O
`
n=0
n
2
r
nn
+2
O
`
n=0
n
r
nn
+
O
`
n=0
r
nn
=
O
`
n=0
n
2
r
nn
+2
O
`
n=0
n
r
nn
+1.
(B4)
Collecting terms, we have
O
`
n=0
n
r
?
nn
= [
g
(n
_
+1)
g
(2n
_
+1)+
g
n
_
]
O
`
n=0
n
2
r
nn
+[
g
(n
_
+1)
g
n
_
+2
g
n
_
]
O
`
n=0
n
r
nn
+
g
n
_
=–
g
O
`
n=0
n
r
nn
+
g
n
_
,
(B5)
or
dkn
ˆ
l
dt
=–
g
kn
ˆ
l+
g
n
_
. (B6)
In the steady state, kn
ˆ
l = n
_
, independent of
g
(see also,
Ref. [121], p. 97).
Acknowledgments
We gratefully acknowledge the support of the
National Security Agency, the Army Research Office,
and the Office of Naval Research. We thank C. Myatt
and C. Wood for comments and suggestions on the
manuscript. We acknowledge useful discussions with
P. Bardroff, R. Blatt, I. Cirac, T. Darling, L.
Davidovich, A. Despain, D. DiVincenzo, A. Ekert,
B. Esry, N. Gisin, S. Haroche, M. Holland,
M. Holzscheiter, R. Hughes, D. James, J. Kimble,
P. Knight, S. Lloyd, G. Milburn, J. Preskill, W.
Schleich, A. Steane, W. Vogel, P. Zoller, and W. Zurek.
10. References
[1] J. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995).
[2] C. A. Blockley, D. F. Walls, and H. Risken, Europhys. Lett. 17,
509 (1992).
[3] D. J. Wineland, J. J. Bollinger, W. M. Itano, F. L. Moore, and
D. J. Heinzen, Phys. Rev. A46, R6797 (1992).
[4] J. I. Cirac, A. S. Parkins, R. Blatt, and P. Zoller, Phys. Rev.
Lett. 70, 556 (1993).
[5] J. I. Cirac, R. Blatt, A. S. Parkins, and P. Zoller, Phys. Rev.
Lett. 70, 762 (1993).
[6] H. Zeng, and F. Lin, Phys. Rev. A48, 2393 (1993).
[7] J. I. Cirac, R. Blatt, A. S. Parkins, and P. Zoller, Phys. Rev.
A49, 1202 (1994).
[8] J. I. Cirac, R. Blatt, and P. Zoller, Phys. Rev. A49, R3174
(1994).
[9] D. J. Wineland, J. J. Bollinger, W. M. Itano, and D. J. Heinzen,
Phys. Rev. A50, 67 (1994).
[10] R. Dum, P. Marte, T. Pellizzari, and P. Zoller, Phys. Rev. Lett.
73, 2829 (1994).
[11] H. Zeng, and F. Lin, Phys. Rev. A50, R3589 (1994).
[12] J. Eschner, B. Appasamy, and P. E. Toschek, Phys. Rev. Lett.
74, 2435 (1995).
[13] R. Blatt, J. I. Cirac, and P. Zoller, Phys. Rev. A52, 518 (1995).
[14] H. Zeng and F. Lin, Phys. Rev. A52, 809 (1995).
[15] S. Wallentowitz and W. Vogel, Phys. Rev. Lett. 75, 2932
(1995).
[16] W. Vogel and R. L. de Matos Filho, Phys. Rev. A52, 4214
(1995).
[17] C. Monroe, D. M. Meekhof, B. E. King, W. M. Itano, and
D. J. Wineland, Phys. Rev. Lett. 75, 4714 (1995).
[18] G. Schrade, V. I. Man’ko, W. P. Schleich, and R. J. Glauber,
Quantum Semiclassical Opt. 7, 307 (1995).
[19] R. L. de Matos Filho and W. Vogel, Phys. Rev. Lett. 76, 608
(1996).
[20] R. L. de Matos Filho and W. Vogel, Phys. Rev. Lett. 76, 4520
(1996).
323
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[21] D. M. Meekhof, C. Monroe, B. E. King, W. M. Itano, and D.
J. Wineland, Phys. Rev. Lett. 76, 1796 (1996); erratum, Phys.
Rev. Lett. 77, 2346 (1996).
[22] J. F. Poyatos, R. Walser, J. I. Cirac, P. Zoller, Phys. Rev. A 53,
R1966 (1996).
[23] S.-C. Gou and P. L. Knight, Phys. Rev. A54, 1682 (1996).
[24] C. D’Helon and G. J. Milburn, Phys. Rev. A54, R25 (1996).
[25] C. D’Helon and G. J Milburn, Phys. Rev. A54, 5141 (1996).
[26] J. F. Poyatos, J. I. Cirac, R. Blatt, and P. Zoller, Phys. Rev. A54,
1532 (1996).
[27] S. Wallentowitz and W. Vogel, Phys. Rev. A 54, 3322 (1996).
[28] S. Wallentowitz and W. Vogel, Phys. Rev. A55, 4438 (1997).
[29] S.-C. Gou, J. Steinbach, and P. L. Knight, Phys. Rev. A 54,
R1014 (1996).
[30] S.-C. Gou, J. Steinbach, and P. L. Knight, Phys. Rev. A 54, 4315
(1996).
[31] S.-C. Gou, J. Steinbach, and P. L. Knight, Phys. Rev. A55, 3719
(1997).
[32] L. Davidovich, M. Orszag, and N. Zagury, Phys. Rev. A54,
5118 (1996).
[33] M. M. Nieto, Phys. Lett. A 219, 180 (1996).
[34] P. J. Bardroff, C. Leichtle, G. Schrade, and W. P. Schleich, Phys.
Rev. Lett. 77, 2198 (1996).
[35] P. J. Bardroff, C. Leichtle, G. Schrade, and W. P. Schleich, Act.
Phys. Slov. 46, 1 (1996).
[36] J. C. Retamal, and N. Zagury, Phys. Rev. A55, 2387 (1997).
[37] C. C. Gerry, S.-C. Gou, and J. Steinbach, Phys. Rev. A55, 630
(1997).
[38] C. C. Gerry, Phys. Rev. A55, 2478 (1997).
[39] S. Wallentowitz, R. L. de Matos Filho, and W. Vogel, Phys. Rev.
A56, 1205 (1997).
[40] J. I. Cirac, A. S. Parkins, R. Blatt, P. Zoller, Adv. At. Mol. Phys.
37, 237 (1996).
[41] C. S. Adams, M. Sigel, and J. Mlynek, Physics Reports 240,
140 (1994).
[42] P. Berman, ed. , Atom Interferometry, Suppl. 3, Adv. At. Mol.
Phys. (Academic, San Diego, CA, 1997).
[43] P. R. Berman, ed. , Cavity Quantum Electrodynamics, (Aca-
demic, Boston, MA, 1994).
[44] F. Diedrich, J. C. Bergquist, W. M. Itano, and D. J. Wineland,
Phys. Rev. Lett. 62, 403 (1989).
[45] C. Monroe, D. M. Meekhof, B. E. King, S. R. Jefferts, W. M.
Itano, D. J. Wineland, and P. Gould, Phys. Rev. Lett. 75, 4011
(1995).
[46] D. J. Wineland, J. C. Bergquist, J. J. Bollinger, and W. M. Itano,
Physica Scripta, T59, 286. (1995).
[47] C. Monroe, D. M. Meekhof, B. E. King, and D. J. Wineland,
Science 272, 1131 (1996).
[48] D. J. Heinzen, and D. J. Wineland, Phys. Rev. A42, 2977
(1990).
[49] P. S. Jessen and I. H. Deutsch, in Adv. At. Mol. Opt. Phys. 37,
B. Bederson and H. Walther, eds., Academic Press, (1996) p.
95.
[50] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman,
and E. A. Cornell, Science 269, 198 (1995).
[51] C. C. Bradley, C. A. Sackett, J. J. Tollett, and R. G. Hulet, Phys.
Rev. Lett. 75, 1687 (1995).
[52] C. C. Bradley, C. C. , C. A. Sackett, and R. G. Hulet, Phys. Rev.
Lett. 78, 985 (1997).
[53] K. B. Davis, M.-O. Mewes, M. R. Andrews, N. J. van Druten,
D. S. Durfee, D. M. Kurn, W. Ketterle, Phys. Rev. Lett. 75,
3969 (1995).
[54] D. M. Greenberger, M. A. Horne, and A. Zeilinger, Phys. Today
46 (8), 22 (1993).
[55] W. H. Zurek, Progress of Theoretical Physics 89, 281 (1993).
[56] W. H. Zurek, Vistas in Astronomy, 37, 185 (1993).
[57] W. H. Zurek, Physics Today 44, 36 (1991).
[58] A. Garg, Phys. Rev. Lett. 77, 964 (1996).
[59] A. Garg, preprint quant-ph/9710053.
[60] A. Steane, Appl. Phys. B 64, 623 (1997).
[61] D. F. James, Appl. Phys. B 66, 181 (1998).
[62] D. F. James, R. J. Hughes, E. H. Knill, R. Laflamme, and A. G.
Petschek, in Photonic Quantum Computing, S. P. Hotaling and
A. R. Pirich, eds., Proc. SPIE, 3076, (1997) p. 42.
[63] R. J. Hughes, D. F. V. James, E. H. Knill, R. Laflamme, and A.
G. Petschek, Phys. Rev. Lett. 77, 3240 (1996).
[64] R. J. Hughes, D. F. V. James, J. J. Gomez, M. S. Gulley, M. H.
Holzscheiter, P. G. Kwiat, S. K. Lamoreaux, C. G. Peterson,
V. D. Sandberg, M. M. Schauer, C. M. Simmons, C. E.
Thorburn, D. Tupa, P. Z. Wang, and A. G. White, Fortsch.
Phys. 46 (4-6), 329 (1998).
[65] D. J. Wineland, C. Monroe, D. M. Meekhof, B. E. King, and D.
Leibfried, W. M. Itano, J. C. Bergquist, D. Berkeland, J. J.
Bollinger, and J. Miller, in Proc. 15th Int. Conf. on Atomic
Physics, Amsterdam, H. B. van Linden van den Heuvell, J. T.
M. Walraven, and M. W. Reynolds, eds., in World Scientific,
Singapore (1997) p. 31.
[66] D. J. Wineland, C. Monroe, W. M. Itano, B. E. King, D.
Leibfried, D. M. Meekhof, C. Myatt, and C. Wood, Forsch.
Phys. 46 (4-6), 363 (1998).
[67] H. G. Dehmelt, Adv. At. Mol. Phys. 3, 53 (1967).
[68] H. G. Dehmelt, Adv. At. Mol. Phys. 5, 109 (1969).
[69] D. J. Wineland, W. M. Itano, and R. S. VanDyck, Jr. , Adv. At.
Mol. Phys. 19, 135 (1983).
[70] R. C. Thompson, Adv. At. Mol. Phys. 31, 253 (1993).
[71] P. K. Ghosh, Ion Traps (Clarendon, Oxford, 1995).
[72] M. G. Raizen, J. M. Gilligan, J. C. Bergquist, W. M. Itano, and
D. J. Wineland, Phys. Rev. A45, 6493(1992).
[73] J. Drees and W. Paul, Z. Phys. 180, 340 (1964).
[74] M. E. Poitzsch, J. C. Bergquist, W. M. Itano, and D. J.
Wineland, Rev. Sci. Instrum. 67, 129 (1996).
[75] J. J. Bollinger, J. D. Prestage, W. M. Itano, and D. J. Wineland,
Phys. Rev. Lett. 54, 1000 (1985).
[76] J. J. Bollinger, D. J. Heinzen, W. M. Itano, S. L. Gilbert, and
D. J. Wineland, IEEE Trans. Instrum. Measurement 40, 126
(1991).
[77] R. S. Van Dyck Jr. , D. L. Farnham, and P. B. Schwinberg,
Phys. Scrip. T59, 134 (1995).
[78] F. DiFilippo, V. Natarajan, M. Bradley, F. Palmer, and D. E.
Pritchard, Physica Scripta T59, 144 (1995).
[79] G. Gabrielse, D. Phillips, W. Quint, H. Kalinowsky, G.
Rouleau, and W. Jhe, Phys. Rev. Lett. 74, 3544 (1995).
[80] H. Totsuji and J.-L. Barrat, Phys. Rev. Lett. 60, 2484 (1988).
[81] D. H. E. Dubin, Phys. Rev. Lett. 71, 2753 (1993).
[82] J. P. Schiffer, Phys. Rev. Lett. 70, 818 (1993).
[83] T. P. Meyrath and D. F. V. James, Phys. Lett. A 240, 37 (1998).
[84] B. Esry and B. Paul, 1997, unpublished.
[85] R. J. Cook, D. G. Shankland, and A. L. Wells, Phys. Rev. A31,
564 (1985).
[86] M. Combescure, Ann. Inst. Henri Poincare´ 44, 293 (1986).
[87] L. S. Brown, Phys. Rev. Lett. 66, 527 (1991).
[88] R. J. Glauber, in Foundations of Quantum Mechanics, T. D.
Black, M. M. Nieto, H. S. Pillof, M. O. Scully, R. M. Sinclair,
eds., World Scientific, Singapore (1992), p. 23.
[89] R. J. Glauber, in Laser Manipulation of Atoms and Ions,
E. Arimondo, W. D. Phillips, and F. Strumia, eds., North
Holland, Amsterdam (1992), p. 643.
[90] V. N. Gheorghe and F. Vedel, Phys. Rev. A45, 4828 (1992).
324
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[91] B. Baseia, R. Vyas, and V. S. Bagnato, Quant. Opt. 5, 155
(1993).
[92] J. I. Cirac, L. J. Garay, R. Blatt, A. S. Parkins, and P. Zoller,
Phys. Rev. A49, 421 (1994).
[93] R. P. Feynman, F. L. Vernon, Jr. , and R. W. Hellwarth, J. Appl.
Phys. 28, 49 (1957).
[94] C. Cohen-Tannoudji, B. Diu, F. Laloe, Quantum Mechanics,
John Wiley and Sons, New York (1977) p. 405.
[95] W. Nagourney, J. Sandberg, and H. G. Dehmelt, Phys. Rev.
Lett. 56, 2797 (1986).
[96] Th. Sauter, R. Blatt, W. Neuhauser, and P. E. Toschek, Phys.
Rev. Lett. 57, 1696 (1986).
[97] J. C. Bergquist, R. G. Hulet, W. M. Itano, and D. J. Wineland,
Phys. Rev. Lett. 57, 1699 (1986).
[98] R. Blatt and P. Zoller, Eur. J. Phys. 9, 250 (1988).
[99] R. J. Cook, in Progress in Optics, XXVIII, E. Wolf, eds., North
Holland, Amsterdam (1990) p. 361.
[100] W. M. Itano, J. C. Bergquist, J. J. Bollinger, J. M. Gilligan,
D. J. Heinzen, F. L. Moore, M. G. Raizen, and D. J. Wineland,
Phys. Rev. A47, 3554 (1993).
[101] D. J. Wineland, J. C. Bergquist, Wayne M. Itano, and R. E.
Drullinger, Opt. Lett. 5, 245 (1980).
[102] H. G. Dehmelt and F. L. Walls, Phys. Rev. Lett. 21, 127 (1968).
[103] D. J. Wineland and H. G. Dehmelt, J. Appl. Phys. 46, 919
(1975).
[104] M. N. Gaboriaud, M. Desaintfuscien, and F. G. Major, Int. J.
Mass Spectrom. Ion Phys. 41, 109 (1981).
[105] K. E. Kahill and R. J. Glauber, Phys. Rev. 177, 1857 (1969).
[106] D. J. Wineland and W. M. Itano, Phys. Rev. A20, 1521 (1979).
[107] E. T. Jaynes and F. W. Cummings, Proc. IEEE 51, 89 (1963).
[108] H. G. Dehmelt, Science 247, 539 (1990).
[109] H. Goldstein, Classical Mechanics, Addison-Wesley, Reading,
MA (1950).
[110] J. Steinbach, J. Twamley, and P. L. Knight, Phys. Rev. A56,
4815 (1997).
[111] J. E. Thomas, P. R. Hemmer, S. Ezekiel, C. C. Leiby, R. H.
Picard, and C. R. Willis, Phys. Rev. Lett. 48, 867 (1982).
[112] W. Happer, Rev. Mod. Phys. 44, 169 (1970).
[113] D. J. Wineland and H. G. Dehmelt, Bull. Am. Phys. Soc. 20,
637 (1975).
[114] D. J. Wineland, R. E. Drullinger, and F. L. Walls, Phys. Rev.
Lett. 40, 1639 (1978).
[115] W. Neuhauser, M. Hohenstatt, P. Toschek, and H. Dehmelt,
Phys. Rev. Lett 41, 233 (1978).
[116] D. J. Wineland, and W. M. Itano, Physics Today 40, (6), p. 34
(1987).
[117] B. Appasamy, I. Siemers, Y. Stalgies, J. Eschner, R. Blatt, W.
Neuhauser, and P. E. Toschek, Appl. Phys. B60, 473 (1995).
[118] G. Morigi, J. I. Cirac, M. Lewenstein, and P. Zoller, Europhys.
Lett. 39, 13 (1997).
[119] V. B. Braginsky and F. Ya. Khalili, Quantum Measurements,
Cambridge University Press, Cambridge (1991).
[120] M. F. Bocko and R. Onofrio, Rev. Mod. Phys. 68, 755 (1996).
[121] D. F. Walls and G. J. Milburn, Quantum Optics, (Springer-
Verlag, Berlin/Heidelberg, 1994).
[122] J. P. Gordon and A. Ashkin, Phys. Rev. Lett. 21, 1606 (1980).
[123] J. Dalibard and C. Cohen-Tannoudji, J. Opt. Soc. Am. B2, 1707
(1985).
[124] H. J. Hope and C. M. Savage, Phys. Rev. A53, 1697 (1996).
[125] D. J. Wineland, J. C. Bergquist, J. J. Bollinger, W. M. Itano, F.
L. Moore, J. M. Gilligan, M. G. Raizen, D. J. Heinzen, C. S.
Weimer, and C. H. Manney, in Laser Manipulation of Atoms
and Ions, E. Arimondo, W. D. Phillips, and F. Strumia, eds.,
North-Holland, Amsterdam (1992), p. 553.
[126] R. L. de Matos Filho and W. Vogel, Phys. Rev. A54, 4560
(1996).
[127] C. K. Law and J. H. Eberly, Phys. Rev. Lett. 76, 1055 (1996).
[128] C. K. Law, J. H. Eberly, and B. Kneer, J. Mod. Opt. 44, 2149
(1997).
[129] S. A. Gardiner, J. I. Cirac, and P. Zoller, Phys. Rev. A55, 1683
(1997).
[130] B. Kneer and C. K. Law, Phys. Rev. A 57, 2096 (1998).
[131] D. Leibfried, D. M. Meekhof, C. Monroe, B. E. King, W. M.
Itano, and D. J. Wineland, Phys. Rev. Lett. 77, 4281 (1996).
[132] D. Leibfried, D. M. Meekhof, C. Monroe, B. E. King, W. M.
Itano, and D. J. Wineland, J. Mod. Opt. 44, 2485 (1997); D.
Leibfried, T. Pfau, and C. Monroe, Phys. Today 51 (4), 22
(1998).
[133] Ch. Kurtsiefer, T. Pfau, and J. Mlynek, Nature 386, 150 (1997).
[134] P. Benioff, J. Stat. Phys. 22, 563 (1980).
[135] P. Benioff, Phys. Rev. Lett. 48, 1581 (1982).
[136] R. Feynman, Int. J. Theor. Phys. 21, 467 (1982).
[137] R. Feynman, Opt. News 11, 11 (1985).
[138] D. Deutsch, Proc. R. Soc. Lond. A400, 97 (1985).
[139] P. Shor, Proc. 35th Ann. Symp. Found. Comp. Sci., IEEE
Computer Society Press, New York (1994) p. 124.
[140] A. Ekert and R. Jozsa, Rev. Mod. Phys. 68, 733 (1996).
[141] A. Barenco, Contemporary Phys. 37, 375 (1996).
[142] A. Steane, Rep. Prog. Phys. 61, 117 (1998).
[143] S. Lloyd, Science 261, 1569 (1993).
[144] S. Lloyd, Science 273, 1073 (1996).
[145] L. K. Grover, Phys. Rev. Lett. 79, 325 (1997).
[146] C. Zalka, Proc. Roy. Soc. Lond., Math. Phys. Sci. (UK) A454,
313 (1998).
[147] S. Wiesner, quant-ph/9603028.
[148] D. S. Abrams and Lloyd, S. , Phys. Rev. Lett. 79, 2586 (1997).
[149] B. M. Boghosian and W. Taylor, Int. J. Mod. Phys. (Singapore)
C8, 705 (1998).
[150] D. A. Lidar and O. Biham, Phys. Rev. E56, 3661 (1997).
[151] S. J. van Enk, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. 78, 429
(1997).
[152] S. J. van Enk, J. I. Cirac, and P. Zoller, Science 279, 205 (1998).
[153] W. G. Teich, K. Obermayer, and G. Mahler, Phys. Rev. B37,
8111 (1988).
[154] G. P. Berman, G. D. Doolen, D. D. Holm, and V. I. Tsifrinovich,
Phys. Lett. 193, 444 (1994).
[155] D. P. DiVincenzo, Phys. Rev. A51, 1015 (1995) and references
therein.
[156] D. P. DiVincenzo, Science 270, 255 (1995).
[157] A. Barenco, D. Deutsch, A. Ekert, and R. Jozsa, Phys. Rev.
Lett. 74, 4083 (1995).
[158] T. Sleator and H. Weinfurter, Phys. Rev. Lett. 74, 4087 (1995).
[159] T. Pellizzari, S. A. Gardiner, J. I. Cirac, and P. Zoller, Phys.
Rev. Lett. 75, 3788 (1995).
[160] P. Domokos, J. M. Raimond, M. Brune, and S. Haroche, Phys.
Rev. A52, 3554 (1995).
[161] P. Domokos, J. M. Raimond, M. Brune, and S. Haroche, in
Laser Spectroscopy, XII International Conf. , Capri, Italy, June
1995, M. Inguscio, M. Allegrini, and A. Sasso, eds., World
Scientific, Singapore (1996) p. 69.
[162] Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, and H. J.
Kimble, Phys. Rev. Lett. 75, 4710 (1995).
[163] W. Lange, Q. A. Turchette, C. J. Hood, H. Mabuchi, and H. J.
Kimble, in Microcavities and Photonic Bandgaps, J. Rarity and
C. Weisbuch, eds., Kluwer (1996) p. 443.
[164] R. To¨rma¨ and S. Stenholm, Phys. Rev. A54, 4701 (1996).
[165] N. A. Gershenfeld and I. L. Chuang, Science 275, 350 (1997).
[166] D. G. Cory, A. F. Fahmy, and T. F. Havel, Proc. Nat. Acad.
Sci.USA 94, 1634 (1997).
[167] V. Privman, I. D. Vagner, and G. Kventsel, Phys. Lett. A239,
141 (1998).
325
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[168] D. Loss and D. P. DiVincenzo, Phys. Rev. A57, 120 (1998).
[169] M. F. Bocko, A. M. Herr, and M. J. Feldman, IEEE Trans.
Appl. Supercond. (USA) 7, 3638 (1997).
[170] A. Barenco, Proc. R. Soc. Lond. A449, 697 (1995).
[171] A. Barenco, C. H. Bennett, R. Cleve, D. P. DiVincenzo, N.
Margolus, P. Shor, T. Sleator, J. Smolin, and H. Weinfurter,
Phys. Rev. A52, 3457 (1995).
[172] M. Brune, P. Nussenzveig, F. Schmidt-Kaler, F. Bernardot, A.
Maali, J. M. Raimond, and S. Haroche, Phys. Rev. Lett. 72,
3339 (1994).
[173] N. F. Ramsey, Molecular Beams, Oxford University, London
(1963).
[174] C. Monroe, D. Leibfried, B. E. King, D. M. Meekhof, W. M.
Itano, and D. J. Wineland, Phys. Rev. A55, R2489 (1997).
[175] P. W. Shor, Phys. Rev. A52, R2493 (1995).
[176] A. Steane, Phys. Rev. Lett. 77, 793 (1996); Phys. Rev. A54,
4741 (1996).
[177] A. Steane, Proc. Roy. Soc. Lond. A452, 2551 (1996).
[178] A. R. Calderbank and P. W. Shor, Phys. Rev. A54, 1098 (1996).
[179] R. LaFlamme, C. Miquel, J. P. Paz, and WH. Zurek, Phys. Rev.
Lett. 77, 198 (1996).
[180] A. Ekert and C. Macchiavello, Phys. Rev. Lett. 77, 2585 (1996).
[181] C. H. Bennett, D. DiVincenzo, J. A. Smolin, and W. K. Woot-
ers, Phys. Rev. A54, 3825 (1996).
[182] J. I. Cirac, T. Pellizzari, and P. Zoller, Science 273, 1207
(1996).
[183] D. P. DiVincenzo and P. W. Shor, Phys. Rev. Lett. 77, 3260
(1996).
[184] E. Knill and R. Laflamme, Phys. Rev. A55, 900 (1997).
[185] E. Knill, R. Laflamme, and W. Zurek, Science 279, 342 (1998).
[186] A. Yu Kitaev, Russian Math. Surveys, 1997, to be published.
[187] C. D’Helon and G. J. Milburn, Phys. Rev. A56, 640 (1997).
[188] M. B. Plenio and P. L. Knight, Proc. R. Soc. Lond. A453, 2017
(1997).
[189] D. J. Wineland, J. C. Bergquist, J. J. Bollinger, W. M. Itano, D.
J. Heinzen, S. L. Gilbert, C. H. Manney, and M. G. Raizen,
IEEE Trans. Ultrasonics Ferroelectrics Frequency Control 37,
515 (1990).
[190] D. J. Wineland, J. C. Bergquist, D. J. Berkeland, J. J. Bollinger,
F. C. Cruz, W. M. Itano, B. M. Jelenkovic´, B. E. King, D. M.
Meekhof, J. D. Miller, C. Monroe, and J. N. Tan, in Proc. Fifth
Symp. Freq. Standards and Metrology, J. C. Bergquist, ed.,
World Scientific, Singapore (1996) p. 11.
[191] D. J. Berkeland, J. D. Miller, J. C. Bergquist, W. M. Itano, and
D. J. Wineland, Phys. Rev. Lett. 80, 2089 (1998).
[192] B. Yurke, Phys. Rev. Lett. 56, 1515 (1986).
[193] B. Yurke, S. L. McCall, and J. R. Klauder, Phys. Rev. A33,
4033 (1986).
[194] J. J. Bollinger, D. J. Wineland, W. M. Itano, and D. J. Heinzen,
Phys. Rev. A54, R4649 (1996).
[195] M. J. Holland and K. Burnett, Phys. Rev. Lett. 71, 1355 (1993).
[196] T. Kim, O. Pfister, M. J. Holland, J. Noh, and J. L. Hall, Phys.
Rev. A57, 4004 (1998).
[197] P. Bouyer and M. Kasevich, Phys. Rev. A56, R1083 (1997).
[198] S. F. Huelga, C. Macchiavello, T. Pellizzari, A. K. Ekert, M. B.
Plenio, and J. I. Cirac, Phys. Rev. Lett. 79, 3865 (1997).
[199] J. R. Anglin, J. P. Paz, and W. H. Zurek, Phys. Rev. A55, 4041
(1997).
[200] J. R. Anglin and W. H. Zurek, in Proceedings of the XVI
Moriond Workshop, R. Ansari, Y. , Giraud-He´raud, and J. Traˆn
Thanh Vaˆn, eds., Editions Frontieres (1996) p. 263.
[201] J. A. Wheeler and W. H. Zurek, eds., Quantum theory and
Measurement, Princeton University, Princeton (1983).
[202] D. F. Walls and G. J. Milburn, Phys. Rev. A31, 2403 (1985).
[203] C. M. Savage and D. F. Walls, Phys. Rev. A32, 2316 (1985).
[204] M. J. Collett, Phys. Rev. A38, 2233 (1988).
[205] M. S. Kim and V. Buzˆek, Phys. Rev. A46, 4239 (1992).
[206] P. Goetsch, R. Graham, and F. Haake, Phys. Rev. A51, 3112
(1995).
[207] M. Brune, E. Hagley, J. Dreyer, X. Maiˆtre, A. Maali, C.
Wunderlich, J. M. Raimond, and S. Haroche, Phys. Rev. Lett.
77, 4887 (1996).
[208] G. S. Agarwal, Phys. Rev. A4, 739 (1971).
[209] C. W. Gardiner, Handbook of Stochastic Methods, Springer-
Verlag, Berlin (1983).
[210] D. J. Daniel and G. J. Milburn, Phys. Rev. A39, 4628 (1989).
[211] S. R. Jefferts, C. Monroe, E. W. Bell, and D. J. Wineland, Phys.
Rev. A51, 3112 (1995).
[212] W. Shockley, J. Appl. Phys. 9, 635 (1938).
[213] M. A. Phillip, F. Gelbard, and S. Arnold, J. Coll. Int. Sci. 91,
507 (1983).
[214] G. Gabrielse, Phys. Rev. A29, 462 (1984).
[215] E. C. Beaty, Phys. Rev A33, 3645 (1986).
[216] T. H. Boyer, Phys. Rev. A9, 68 (1974).
[217] S. K. Lamoreaux, Phys. Rev. A56, 4970 (1997).
[218] T. A. Savard, K. M. O’Hara, and J. E. Thomas, Phys. Rev. A 56,
R1095 (1997).
[219] H. Walther, Adv. At. Mol. Phys. 31, 137 (1993); J. W. Emmert,
M. Moore, and R. Blu¨mel, Phys. Rev. A48, R1757 (1993).
[220] C. S. Weimer, F. L. Moore, and D. J. Wineland, Phys. Rev. Lett.
70, 2553 (1993).
[221] R. Alheit, Th. Gudjons, S. Kleineidam, and G. Werth, Rapid
Commun. Mass Spectrometry 10, 583 (1996).
[222] Y. Wang, J. Franzen, and K. P. Wanczek, Int. J. Mass Spec-
trometry Ion Processes 124, 125 (1993).
[223] J. B. Camp, T. W. Darling, and R. E. Brown, J. Appl. Phys. 69,
7126 (1991).
[224] N. Yu, W. Nagourney, and H. Dehmelt, J. Appl. Phys. 69, 3779
(1991).
[225] R. G. DeVoe and R. G. Brewer, in Laser Spectroscopy, M.
Inguscio, M. Allegrini, and A. Sasso, eds., World Scientific,
Sangapore (1996) p. 175.
[226] G. Janik, W. Nagourney, and H. Dehmelt, J. Opt. Soc. Am. B2,
1251 (1985).
[227] J. T. Ho¨ffges, H. W. Baldauf, T. Eichler, S. R. Helmfrid, and H.
Walther, Opt. Commun. 133, 170 (1997).
[228] D. Berkeland, J. D. Miller, J. C. Bergquist, W. M. Itano, and D.
J. Wineland, J. Appl. Phys. 83, 5025 (1998).
[229] Ch. Kleint, Ann. Phys. 10, 309 (1963).
[230] G. W. Timm and A. Van der Ziel, Physica 32, 1333 (1966).
[231] P. A. Redhead, J. P. Hobson, E. V. Kornelsen, The Physical
Basis of Ultrahigh Vacuum, Chapman and Hall, London
(1968).
[232] L. Schmidt and R. Gomer, J. Chem. Phys. 42, 3573 (1965).
[233] F. C. Tompkins, Chemisorption of Gases on Metals, Academic,
London (1978).
[234] L. D. Landau and E. M. Lifshitz, Mechanics, Pergamon, Ox-
ford/New York (1960).
[235] J. C. Bergquist, W. M. Itano, and D. J. Wineland, in Frontiers
in Laser Spectroscopy, T. W. Ha¨nsch and M. Inguscio, eds.,
North Holland, Amsterdam (1994) p. 359.
[236] K. Sugiyama and J. Yoda, Phys. Rev. A55, R10 (1997).
[237] J. B. Hasted, Physics of Atomic Collisions, Butterworth & Co.,
New York (1972).
[238] L. D. Landau and E. M. Lifshitz, Quantum Mechanics—Non-
relativistic Theory, Pergamon, Oxford/NY (1977).
[239] F. G. Major and H. G. Dehmelt, Phys. Rev. 170, 91 (1968).
[240] L. G. Lutterbach and L. Davidovich, Phys. Rev. Lett. 78, 2547
(1997).
326
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[241] L. Davidovich, M. Brune, J. M. Raimond, and S. Haroche,
Phys. Rev. A53, 1295 (1996).
[242] J. C. Bergquist, ed., Proc. Fifth Symp. Freq. Standards and
Metrology, World Scientific, Singapore (1995).
[243] P. T. H. Fisk, Rep. Prog. Phys. 60, 1 (1997).
[244] P. T. H. Fisk, M. J. Sellars, M. A. Lawn, C. Coles, A. G. Mann,
and D. G. Blair, IEEE Trans. Instrum. Meas. 44, 113 (1995).
[245] M. Zhu and J. L. Hall, J. Opt. Soc. Am. B10, 802 (1993).
[246] J. Hall, Frequency-Stabilized Lasers and Their Applications,
SPIE Proc. Vol. 1837 (1993) p. 2.
[247] M. Zhu and J. L. Hall, Experimental Methods in the Physical
Sciences, F. B. Dunning and R. Hulet, eds., Academic, San
Diego (1997) p. 103.
[248] F. Plumelle, M. Desaintfuscien, J. L. Duchene and C. Audoin,
Opt. Commun. 34, 71 (1980).
[249] A. A. Madej and J. D. Sankey, Phys. Rev. A41, 2621 (1990).
[250] N. Yu, W. Nagourney, H. Dehmelt, Phys. Rev. Lett. 78, 4898
(1997).
[251] D. A. Church, Phys. Reports 228, 253 (1993).
[252] D. J. Wineland, W. M. Itano, J. C. Bergquist, and F. L. Walls,
in Proc. 35th Ann. Symp. Freq. Control, IEEE, Piscataway, NJ
(1981) p. 602.
[253] W. M. Itano, L. L. Lewis, and D. J. Wineland, Phys. Rev. A25,
1233 (1982).
[254] N. Yu, X. Zhao, H. Dehmelt, and W. Nagourney, Phys. Rev.
A50, 2738 (1994).
[255] W. M. Itano, 1997, unpublished.
[256] H. G. Dehmelt, IEEE Trans. Instrum. Meas. 31, 83 (1982).
[257] D. J. Wineland, W. M. Itano, J. C. Bergquist, J. J. Bollinger, and
J. D. Prestage, Ann. Phys. Fr. 10, 737 (1985).
[258] H. Carmichael, An Open Systems Approach to Quantum Op-
tics, Lecture Notes in Physics m18, Springer, Berlin (1993).
[259] M. B. Plenio and P. L. Knight, Rev. Mod. Phys., 70, 101 (1998).
[260] R. Landauer, in Proceedings of the Drexel-4 Symposium on
Quantum Nonintegrability, D. H. Feng and B-L. Hu, eds., Inter-
national, Boston (1997).
[261] J. F. Poyatos, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. 78, 390
(1997).
[262] I. L. Chuang and M. A. Nielsen, J. Mod. Optics 44, 2455
(1997).
[263] K. Obenland and A. M. Despain, in Proc. 1
st
NASA Int. Conf.
of Quantum Computing and Quantum Information, to be pub-
lished (quant-ph/9804038).
[264] D. Beckman, A. N. Chari, S. Devabhaktuni, and J. Preskill,
Phys. Rev. A54, 1034 (1996).
[265] A. Barenco, A. Ekert, K.-A. Suominen, and P. To¨rma¨, Phys.
Rev. A54, 139 (1996).
[266] C. Miquel, J. P. Paz, and W. H. Zurek, Phys. Rev. Lett. 78, 3971
(1997).
[267] A. Messiah, Quantum Mechanics, Vol. II North-Holland, Am-
sterdam (1962) p. 744.
[268] J. Lawall and M. Prentiss, Phys. Rev. Lett. 72, 993 (1994).
[269] L. S. Goldner, C. Gerz, R. J. C. Spreuw, S. L. Rolston, C. I.
Westbrook, W. D. Phillips, P. Marte, and P. Zoller, Phys. Rev.
Lett. 72, 997 (1994).
[270] M. Weitz, B. C. Young, and S. Chu, Phys. Rev. Lett. 73, 2563
(1994).
[271] M. Weitz, B. C. Young, and S. Chu, Phys. Rev. A50, 2438
(1994).
[272] N. C. Wong and J. L. Hall, J. Opt. Soc. Am. B2, 1527 (1985).
[273] C. C. Harb, M. B. Gray, H.-A. Bachor, R. Schilling, P. Rotten-
gatter, I. Freitag, and H. Welling, IEEE J. Quant. Elec. 30, 2907
(1994).
[274] S. Schneider and G. J. Milburn, Phys. Rev. A57, 3748 (1998).
[275] R. F. C. Vessot, M. E. Levine, E. M. Mattison, E. L. Blomberg,
T. E. Hoffman, G. U. Nystrom, B. F. Farrel, R. Decher, P. B.
Eby, C. R. Baugher, J. W. Watts, E. L. Teuber, and F. D. Wills,
Phys. Rev. Lett. 45, 2081 (1980).
[276] A. E. Siegman, Lasers, University Science Books, Mill Valley,
CA (1986).
[277] S. Chu (1997) private communication.
[278] J. I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi, Phys. Rev.
Lett. 78, 3221 (1997).
[279] H. Frauenfelder, The Mossbauer Effect, Benjamin, New York
(1963).
[280] H. J. Lipkin, Quantum Mechanics, North Holland, Amsterdam
(1973), Chaps. 2-4.
[281] M. A. Kasevich and S. Chu, Phys. Rev. Lett. 69, 1741 (1992).
[282] D. S. Weiss, B. C. Young, and S. Chu, Appl. Phys. B59, 217
(1994).
[283] R. Kosloff, S. A. Rice, P. Gaspard, D. J. Tannor, Chem. Phys.
139, 20 (1989).
[284] S. Shi and H. Rabitz, J. Chem. Phys. 92, 364 (1990).
[285] S. Rice, Science 258, 412 (1992).
[286] W. S. Warren, H. Rabitz, and M. Dahleh, Science 259, 1581
(1993).
[287] J. I. Cirac, R. Blatt, P. Zoller, and W. D. Phillips, Phys. Rev.
A46, 2668 (1992).
[288] R. L. de Matos Filho and W. Vogel, Phys. Rev. A50, R1988
(1994).
[289] M. B. Plenio and P. L. Knight, Phys. Rev. A53, 2986 (1996).
[290] R. A. Cline, J. D. Miller, M. R. Matthews, and D. J. Heinzen,
Optics Lett., 19, 207 (1994).
[291] J. I. Cirac and P. Zoller (1997) private communication.
[292] M. B. Plenio, V. Vedral, and P. L. Knight, Phys. Rev. A55,67
(1997).
[293] H. Harde, H. Lehmitz, J. Hattendorf-Ledwoch, and R. Blatt,
Appl. Phys. B53, 131 (1991).
[294] J. Preskill, Proc. Roy. Soc. Lond., Math, Phys. Eng. Sci. (UK)
A454, 385 (1998).
[295] R. Brewer, R. G. DeVoe, and R. Kallenbach, Phys. Rev. A46,
R6781 (1992).
[296] I. Klaft et al. , Phys. Rev. Lett. 73, 2425 (1994).
[297] J. R. Crespo Lo´pez-Urrutia, P. Beiersdorfer, D. W. Savin, and
K. Widmann, Phys. Rev. Lett. 77, 826 (1996).
[298] R. J. Gould, Astrophys. J. 423, 522 (1994).
[299] A. Einstein, B. Podolsky, N. Rosen, Phys. Rev. 47, 777 (1935).
[300] J. S. Bell, Physics 1, 195 (1964).
[301] J. F. Clauser and A. Shimony, Rep. Prog. Phys. 41, 1881 (1978).
[302] A. Aspect, J. Dalibard, G. Roger, Phys. Rev. Lett. 49, 1804
(1982).
[303] E. S. Fry, T. Walther, and S. Li, Phys. Rev. A52, 4381 (1995).
[304] E. Hagley, X. Maˆtre, G. Nogues, C. Wunderlich, M. Brune, J.
M. Raimond, and S. Haroche, Phys. Rev. Lett. 79, 1 (1997).
[305] D. Greenberger, M. A. Horne, A. Zeilinger, in Bell’s Theorem,
Quantum Theory, and Conceptions of the Universe,
M. Kafatos, ed., Kluwer Academic, Dordrecht (1989).
[306] N. D. Mermin, Am. J. Phys. 58, 731 (1990).
[307] G. C. Ghirardi, A. Rimini, and T. Weber, Phys. Rev. D34, 470
(1986).
[308] P. Pearle, Phys. Rev. A39, 2277 (1989).
[309] S. Wallentowitz, W. Vogel, I. Siemers, and P. E. Toschek, Phys.
Rev. A54, 943 (1996).
[310] G. Drobny´ and B. Hladky´, Acta Physica Slovaka 47, 277
(1997).
[311] B. Buck and C. V. Sukumar, Phys. Lett. 81A, 132 (1981).
[312] P. L. Knight, Phys. Scripta T12, 51 (1986).
[313] K. Banaszek and P. L. Knight, Phys. Rev. A55, 2368 (1997).
327
Volume 103, Number 3, May–June 1998
Journal of Research of the National Institute of Standards and Technology
[314] G. S. Agarwal and J. Banerji, Phys. Rev. A55, R4007 (1997).
[315] C. D’Helon and G. J. Milburn, Phys. Rev. A52, 4755 (1995).
[316] J. J. Slosser and G. J. Milburn, Phys. Rev. Lett. 75, 418 (1995).
[317] F. E. Harrison, A. S. Parkins, M. J. Collett, D. F. Walls, Phys.
Rev. A55, 4412 (1997).
[318] W. K. Lai, V. Buzˇek, and P. L. Knight, Phys. Rev. A43, 6323
(1991).
[319] B. C. Sanders and G. J. Milburn, Phys. Rev. Lett. 75, 2944
(1995).
[320] K. Abdullah, C. Carlberg, E. D. Commins, H. Gould, and S. B.
Ross, Phys. Rev. Lett. 65, 2347 (1990).
[321] M. Charlton, J. Eades, D. Horva´the, R. J. Hughes, and C.
Zimmerman, Phys. Rep. 241, 65 (1994).
[322] W. Quint and G. Gabrielse, Hyperfine Interactions 76, 379
(1993).
[323] D. J. Bate, K. Dholakia, R. C. Thompson, and D. C. Wilson, J.
Mod. Opt. 39, 305 (1992).
[324] M. Roukes (1997) private communication.
[325] A. N. Cleland and M. L. Roukes, Appl. Phys. Lett. 69, 2653
(1996).
[326] A. Steane, quant-ph/9708021.
[327] J. Vanier, M. Teˆtu, and L.-G. Bernier, IEEE Trans. Instrum.
Meas. IM28, 188 (1979).
[328] F. L. Walls, in Frequency Standards and Metrology, Proc. of
Fourth Symposium on Frequency Standards and Metrology,
Ancona, A. DeMarchi, ed., Springer-Verlag, Berlin (1989) p.
145.
[329] J. Dick, J. D. Prestage, C. A. Greenhall, and L. Maleki, in Proc.
21st Ann. Precise Time and Time Interval (PTTI) Applications
and Planning Meeting (1989) p. 487.
[330] C. Cohen-Tannoudji, in Frontiers in Laser Spectroscopy,
R. Balian, S. Haroche, and S. Liberman, eds., North-Holland
(1977), p. 1.
About the authors: David Wineland, Christopher
Monroe, Wayne Itano, and Dawn Meekhof are staff
physicists in the Time and Frequency Division of the
Physics Laboratory at Boulder. Brian King is a gradu-
ate student in physics at the University of Colorado.
Diedrich Leibfried was a guest researcher during
the preparation of the manuscript; he is currently a
physicist at the Institute for Experimental Physics in
Innsbruck, Austria. The National Institute of Standards
and Technology is an agency of the Technology
Administration, U.S. Department of Commerce.
Note added in proof
Some recent experiments are relevant to the subjects
of this paper. In addition to the work at NIST [72,191]
and Garching [219], other groups have realized
Doppler-cooled “arrays” of ions which might be used as
quantum registers [331-335]. The motional modes of
two trapped ions have recently been laser-cooled to the
ground state [336] and differentially addressed to make
approximate Bell states [337]. Recent experiments on
neutral atoms confined in optical lattices have employed
sideband laser cooling using stimulated-Raman transi-
tions [338,339]; these experiments have achieved cool-
ing to kn
ˆ
l=0 with high efficiency [338,340]. Quantum
state engineering and manipulation, and schemes for
quantum logic will follow [340-342].
References added in proof
[331] R. G. DeVoe and R. G. Brewer, Phys. Rev. Lett. 76, 2049
(1996); Ch. Kurtsiefer and R. G. DeVoe, Abstract, Int. Quant.
Elect. Conf. (1998).
[332] H. C. Na¨gerl, W. Bechter, J. Eschner, and R. Blatt, Appl. Phys.
B66, 603 (1998); H. C. Na¨gerl, D. Leibfried, F. Schmidt-Kaler,
J. Eschner, and R. Blatt, Optics Express, to be published.
[333] W. Alt, M. Block, P. Seibert, G. Werth, Phys. Rev. A (Rapid
Comm.), to be published.
[334] M. Drewsen, C. Brodersen, L. Hornekaer, J. S. Hangst, and J.
P. Schiffer, Phys. Rev. Lett., to be published.
[335] T. B. Mitchell, J. J. Bollinger, D. H. E. Dubin, X.-P. Huang, W.
M. Itano, and R. H. Baughman, submitted for publication.
[336] B. E. King, C. S. Wood, C. J. Myatt, Q. A. Turchette, D.
Leibfried, W. M. Itano, C. Monroe, and D. J. Wineland, quant-
ph/9803023.
[337] Q. A. Turchette, C. S. Wood, B. E. King, C. J. Myatt, D.
Leibfried, W. M. Itano, C. Monroe, and D. J. Wineland, quant-
ph/9806012.
[338] S. E. Hamann, D. L. Haycock, G. Klose, P. H. Pax, I. H.
Deutsch, and P. S. Jessen, Phys. Rev. Lett. 80, 4149 (1998).
[339] H. Perrin, A. Kuhn, I. Bouchoule, and C. Salomon, Europhys.
Lett. 42, 395 (1998).
]340] C. Salomon, private communication.
[341] I. H. Deutsch and P. S. Jessen, Phys. Rev. A57, 1972 (1998).
[342] P. Zoller, private communication.
328