REPORTS
Cite as: M. Andres et al., Science
10.1126/science.aah3752 (2016).
The detection and manipulation of individual quantum par-
ticles, such as atoms or photons, is now routinely performed
in many quantum physics experiments (1, 2); however, re-
taining the same control in large-scale systems remains an
outstanding challenge. For example, major efforts are cur-
rently aimed at scaling up ion-trap and superconducting
platforms, where high-fidelity quantum computing opera-
tions have been demonstrated in registers consisting of sev-
eral qubits (3, 4). In contrast, ultracold quantum gases
composed of neutral atoms offer inherently large system
sizes. However, arbitrary single atom control is highly de-
manding and its realization is further limited by the slow
evaporative cooling process necessary to reach quantum
degeneracy. Only in recent years has individual particle de-
tection (5, 6) and basic single-spin control (7) been demon-
strated in low entropy optical lattice systems.
Here we demonstrate atom-by-atom assembly of large
defect-free 1D arrays of cold neutral atoms (8, 9).
We use optical microtraps to directly extract individual
atoms from a laser-cooled cloud (1012) and employ recently
demonstrated trapping techniques (1316) and single-atom
position control (1720) to create desired atomic configura-
tions. Central to our approach is the use of single-atom de-
tection and real-time feedback (17, 20) to eliminate the
entropy associated with the probabilistic trap loading (10)
[currently limited to ninety percent loading probability even
with advanced techniques (2123)]. Related to the funda-
mental concept of “Maxwell’s demon” (8, 9), this method
allows us to rapidly create large defect-free arrays, and
when supplemented with appropriate atom-atom interac-
tions (15, 16, 2430) provides a potential platform for scala-
ble experiments with individually controlled neutral atoms.
The experimental protocol is illustrated in Fig. 1A. An ar-
ray of 100 tightly focused optical tweezers is loaded from a
laser-cooled cloud. The collisional blockade effect ensures
that each individual tweezer is either empty or occupied by
a single atom (10). A first high-resolution image yields sin-
gle-atom resolved information about the trap occupation,
which we use to identify empty traps and to switch them
off. The remaining occupied traps are rearranged into a
regular, defect-free array and we detect the final atom con-
figuration with a second high-resolution image.
Our implementation relies on fast, real-time control of
the tweezer positions (Fig. 1B), which we achieve by employ-
ing an acousto-optic deflector (AOD) that we drive with a
multi-tone radio-frequency (RF) signal.
This generates an array of deflected beams, each con-
trolled by its own RF-tone (15, 16). The resulting beam array
is then focused into our vacuum chamber and forms an ar-
ray of optical tweezers, each with a Gaussian waist of
900 nm
, a wavelength of 809 nm, and a trap depth of
/ 0.9 mK
B
Uk
(k
B
, Boltzmann constant) that is homoge-
neous across the array within 2% (31).
The tweezer array is loaded from a laser-cooled cloud of
Rubidium-87 atoms in a magneto-optical trap (MOT). After
Atom-by-atom assembly of defect-free one-dimensional
cold atom arrays
Manuel Endres,
1,2
*
Hannes Bernien,
1
* Alexander Keesling,
1
* Harry Levine,
1
* Eric R. Anschuetz,
1
Alexandre Krajenbrink,
1‡
Crystal Senko,
1
Vladan Vuletic,
3
Markus Greiner,
1
Mikhail D. Lukin
1
1
Department of Physics, Harvard University, Cambridge, MA 02138, USA.
2
Division of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena,
CA 91125, USA.
3
Department of Physics and Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA.
*These authors contributed equally to this work.
Corresponding author. Email: men[email protected]
Present address: CNRS-Laboratoire de Physique Theorique de l’Ecole Normale Superieure, 24 Rue L’Homond, 75231 Paris Cedex, France.
The realization of large-scale fully controllable quantum systems is an exciting frontier in modern physical
science. We use atom-by-atom assembly to implement a platform for the deterministic preparation of
regular one-dimensional arrays of individually controlled cold atoms. In our approach, a measurement and
feedback procedure eliminates the entropy associated with probabilistic trap occupation and results in
defect-free arrays of over 50 atoms in less than 400 milliseconds. The technique is based on fast, real-
time control of 100 optical tweezers, which we use to arrange atoms in desired geometric patterns and to
maintain these configurations by replacing lost atoms with surplus atoms from a reservoir. This bottom-
up approach may enable controlled engineering of scalable many-body systems for quantum information
processing, quantum simulations, and precision measurements.
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 1
on June 7, 2018 http://science.sciencemag.org/Downloaded from
the loading procedure, we let the MOT cloud disperse and
we detect the occupation of the tweezers with fluorescence
imaging. Fast, single-shot, single-atom resolved detection
with 20 ms exposure is enabled by a sub-Doppler laser-
cooling configuration that remains active during the re-
mainder of the sequence (31) (see Fig. 2, A to C). Our fluo-
rescence count statistics show that individual traps are
either empty or occupied by a single atom (10, 31), and we
find probabilistically filled arrays with an average single
atom loading probability of
0.6p
(see Figs. 2A and 3A).
The central part of our scheme involves the rearrange-
ment procedure for assembling defect-free arrays (31) (see
Fig. 1A). In the first step, unoccupied traps are switched off
by setting the corresponding RF-amplitudes to zero. In a
second step, all occupied tweezers are moved to the left un-
til they stack up with the original spacing of 2.6 μm. This
movement is generated by sweeping the RF-tones to change
the deflection angles of the AOD and lasts 3 ms (31). Finally,
we detect the resulting atom configuration with a second
high-resolution image. These steps implement a reduction
of entropy via measurement and feedback. The effect is im-
mediately visible in the images shown in Fig. 2, A and B.
The initial filling of our array is probabilistic, whereas the
rearranged configurations show highly ordered atom arrays.
Our approach also allows us to construct flexible atomic
patterns (Fig. 2C).
The rearrangement procedure creates defect-free arrays
with high fidelity. This can be quantified by considering the
improvement of single atom occupation probabilities (Fig.
3A) and the success probabilities, p
N
, for creating defect-free
arrays of length N (Fig. 3B). The single atom occupation
probability in the left-most forty traps increases from
0.6
before rearrangement to 0.988(3) after rearrangement,
demonstrating our ability for high-fidelity single-atom prep-
aration. Furthermore, the success probabilities for creating
defect-free arrays show an exponential improvement. Prior
to rearrangement, the probability of finding a defect-free
array of length N is exponentially suppressed with p
N
= p
N
where
0.6p
(blue circles, Fig. 3B). After rearrangement,
we find success probabilities as high as p
30
= 0.75(1) and p
50
= 0.53(1) (red circles, Fig. 3B).
The same exponential improvement is observed by con-
sidering the average wait time for producing defect-free ar-
rays, given by T/p
N
, where T = 200 ms is the cycling time of
our experiment (see Fig. 3B). For example, we are able to
generate defect-free arrays of 50 atoms with an average wait
time of less than 400 ms (red circles, Fig. 3C).
The success probabilities can be further enhanced
through multiple repetitions of the rearrangement protocol.
Figure 4 illustrates the procedure in which we target an
atomic array of fixed length and create a reservoir from sur-
plus atoms in a separate zone. After the initial arrangement
of atoms into the target and reservoir zones, we periodically
take images to identify defects in the target array and pull
atoms from the reservoir to fill in these defects. This en-
hances our initial success probabilities at producing defect-
free arrays within one MOT-loading cycle to nearly the ideal
limit (Fig. 4C).
Finally, a similar procedure can be used for correcting
errors associated with atomic loss. This becomes a signifi-
cant limitation for large arrays because for a given lifetime
of an individual atom in the trap τ, the corresponding life-
time of the N atom array scales as τ/N. To counter this loss,
we repeatedly detect the array occupation after longer time
intervals and replenish lost atoms from the reservoir. This
procedure leads to exponentially enhanced lifetimes of our
arrays (Fig. 4D).
These results demonstrate the ability to generate and
control large, defect-free arrays at a fast repetition rate. The
success probabilities are limited by two factors: the initial
number of loaded atoms and losses during rearrangement.
For example, the average total atom number in our array is
59 ± 5 (31), which results in the cutoff in the success proba-
bility in Fig. 3B starting from
50N
(solid line). For short-
er arrays, the fidelity is mostly limited by losses during
rearrangement. These losses are dominated by our finite
vacuum-limited lifetime, which varies from
to
12 s
τ
(depending on the setting of our atomic dispenser
source), and are only minimally increased by the movement
of the atoms (31). The single atom occupation probability is
correspondingly reduced by a factor exp(t
r
/τ), where t
r
=
50 ms is the time for the whole rearrangement procedure
(31). This results in the success probabilities of creating
length-N arrays scaling as exp(t
r
N/τ), which dominates the
slope for N = 50 in Fig. 3B (dashed line). Currently, we reach
vacuum limited lifetimes only with sub-Doppler cooling ap-
plied throughout the sequence. However, the lifetime with-
out cooling could be improved, for example, by using a
different trapping laser and trapping wavelength (31).
The size of the final arrays can be considerably increased
by implementing a number of realistic experimental im-
provements. For example, the initial loading probability
could be enhanced to 0.9 (2123) and the vacuum limited
lifetime could be improved to
60 s
τ
in an upgraded vac-
uum chamber. Increasing the number of traps in the cur-
rent configuration is difficult because of the AOD
bandwidth of
50 MHz
and strong parametric heating in-
troduced when the frequency spacing of neighboring traps
approaches
450 kHz
(31). However, implementing two-
dimensional (2D) arrays could provide a path toward realiz-
ing thousands of traps, ultimately limited by the availability
of laser power and the field of view of high-resolution objec-
tives. Such 2D configurations could be realized by either
directly using a 2D-AOD or by creating a static 2D lattice of
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 2
on June 7, 2018 http://science.sciencemag.org/Downloaded from
traps [using spatial light modulators (14) or optical lattices
(12)] and sorting atoms with an independent AOD (31). With
increased loading efficiencies (2123), realistic estimates for
the rearrangement time t
r
in such 2D arrays indicate that
the robust creation of defect-free arrays of hundreds of at-
oms is feasible (31). Finally, the repetitive interrogation
techniques, in combination with periodic reservoir reload-
ing from a cold atom source (such as a MOT), could be used
to maintain arrays indefinitely.
Atom-by-atom assembly of defect-free arrays forms a
scalable platform with unique possibilities. It combines fea-
tures that are typically associated with ion trapping experi-
ments, such as single-qubit addressability (32, 33) and fast
cycling times, with the flexible optical trapping of neutral
atoms in a scalable fashion. Furthermore, in contrast to sol-
id-state platforms, such atomic arrays are highly homogene-
ous (31) and mostly decoupled from their environment. The
homogeneity of our array should also allow for cooling of
the atomic motion via simultaneous sideband cooling in all
tweezers at once (34, 35).
These features provide an excellent starting point for
multi-qubit experiments, studies of quantum many body
effects and for exploring future applications. The required
interactions between the atoms can be engineered using
several approaches. Even without sideband cooling, exciting
the atoms into high-lying Rydberg states would introduce
strong dipole interactions that can be used for fast entan-
gling gates (24, 25, 27). The parallelism afforded by our flex-
ible atom rearrangement enables efficient diagnostics of
such Rydberg-mediated entanglement. These interactions
may also enable approaches to quantum simulations that
involve both coherent coupling and engineered dissipation
(26, 27), as well as large-scale entangled quantum states for
applications in precision measurements (36).
An alternative approach to engineering interactions in-
volves the integration of atom arrays with nanophotonic
platforms as demonstrated previously (28, 29). These enable
photon-mediated interactions that can be employed to cou-
ple the atoms within a local multi-qubit register or for effi-
cient communication between the registers using a modular
quantum network architecture (3).
Finally, our platform could enable new bottom-up ap-
proaches to studying quantum many-body physics in Hub-
bard models (15, 16, 30), where atomic Mott insulators with
fixed atom number and complex spin patterns could be di-
rectly assembled. This requires atom temperatures close to
the ground state, coherent tunneling between the traps, and
sizable on-site interactions. With side-band cooling, ground
state fractions in excess of 90% have already been demon-
strated (34, 35), and can likely be improved via additional
optical trapping along the longitudinal tweezer axes, which
would also increase on-site interaction strengths. Coherent
tunneling of Rb atoms between similarly sized tweezers has
been observed before by reducing the tweezer distance (15,
16). The parametric heating, currently limiting the minimal
distance between our traps, could be reduced by working
with shallower traps, as needed for tunneling, and by em-
ploying fewer traps to increase the frequency separation
between neighboring traps. Eventually, this approach could
be applied to create ultracold quantum matter composed of
exotic atomic species or complex molecules (37, 38) that are
difficult to cool evaporatively.
REFERENCES AND NOTES
1. S. Haroche, Controlling photons in a box and exploring the quantum to classical
boundary. Ann. Phys.
525
, 753776 (2013). doi:10.1002/andp.201300737
2. D. J. Wineland, Nobel Lecture: Superposition, entanglement, and raising
Schrödinger’s cat. Rev. Mod. Phys.
85
, 11031114 (2013).
doi:10.1103/RevModPhys.85.1103
3. C. Monroe, J. Kim, Scaling the ion trap quantum processor. Science
339
, 1164
1169 (2013).
Medline doi:10.1126/science.1231298
4. M. H. Devoret, R. J. Schoelkopf, Superconducting circuits for quantum
information: An outlook. Science
339
, 11691174 (2013). Medline
doi:10.1126/science.1231930
5. W. S. Bakr, A. Peng, M. E. Tai, R. Ma, J. Simon, J. I. Gillen, S. Fölling, L. Pollet, M.
Greiner, Probing the superfluid-to-Mott insulator transition at the single-atom
level. Science
329
, 547550 (2010). Medline doi:10.1126/science.1192368
6. J. F. Sherson, C. Weitenberg, M. Endres, M. Cheneau, I. Bloch, S. Kuhr, Single-
atom-resolved fluorescence imaging of an atomic Mott insulator. Nature
467
,
6872 (2010).
Medline doi:10.1038/nature09378
7. C. Weitenberg, M. Endres, J. F. Sherson, M. Cheneau, P. Schauss, T. Fukuhara, I.
Bloch, S. Kuhr, Single-spin addressing in an atomic Mott insulator. Nature
471
,
319324 (2011).
Medline doi:10.1038/nature09827
8. D. S. Weiss, J. Vala, A. V. Thapliyal, S. Myrgren, U. Vazirani, K. B. Whaley, Another
way to approach zero entropy for a finite system of atoms. Phys. Rev. A
70
,
040302 (2004).
doi:10.1103/PhysRevA.70.040302
9. J. Vala, A. V. Thapliyal, S. Myrgren, U. Vazirani, D. S. Weiss, K. B. Whaley, Perfect
pattern formation of neutral atoms in an addressable optical lattice. Phys. Rev. A
71
, 032324 (2005). doi:10.1103/PhysRevA.71.032324
10. N. Schlosser, G. Reymond, I. Protsenko, P. Grangier, Sub-poissonian loading of
single atoms in a microscopic dipole trap. Nature
411
, 10241027 (2001).
Medline doi:10.1038/35082512
11. M. Weber, J. Volz, K. Saucke, C. Kurtsiefer, H. Weinfurter, Analysis of a single-
atom dipole trap. Phys. Rev. A
73
, 043406 (2006).
doi:10.1103/PhysRevA.73.043406
12. K. D. Nelson, X. Li, D. S. Weiss, Imaging single atoms in a three-dimensional array.
Nat. Phys.
3
, 556560 (2007). doi:10.1038/nphys645
13. M. J. Piotrowicz, M. Lichtman, K. Maller, G. Li, S. Zhang, L. Isenhower, M.
Saffman, Two-dimensional lattice of blue-detuned atom traps using a projected
Gaussian beam array. Phys. Rev. A
88
, 013420 (2013).
doi:10.1103/PhysRevA.88.013420
14. F. Nogrette, H. Labuhn, S. Ravets, D. Barredo, L. Béguin, A. Vernier, T. Lahaye, A.
Browaeys, Single-atom trapping in holographic 2D arrays of microtraps with
arbitrary geometries. Phys. Rev. X
4
, 021034 (2014);
http://journals.aps.org/prx/abstract/10.1103/PhysRevX.4.021034.
15. A. M. Kaufman, B. J. Lester, C. M. Reynolds, M. L. Wall, M. Foss-Feig, K. R.
Hazzard, A. M. Rey, C. A. Regal, Two-particle quantum interference in tunnel-
coupled optical tweezers. Science
345
, 306309 (2014). Medline
doi:10.1126/science.1250057
16. A. M. Kaufman, B. J. Lester, M. Foss-Feig, M. L. Wall, A. M. Rey, C. A. Regal,
Entangling two transportable neutral atoms via local spin exchange. Nature
527
,
208211 (2015).
Medline doi:10.1038/nature16073
17. Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Förster, M. Khudaverdyan, D.
Meschede, D. Schrader, A. Rauschenbeutel, Quantum engineering: An atom-
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 3
on June 7, 2018 http://science.sciencemag.org/Downloaded from
sorting machine. Nature
442
, 151 (2006). Medline doi:10.1038/442151a
18. J. Beugnon, C. Tuchendler, H. Marion, A. Gaëtan, Y. Miroshnychenko, Y. R. P.
Sortais, A. M. Lance, M. P. A. Jones, G. Messin, A. Browaeys, P. Grangier, Two-
dimensional transport and transfer of a single atomic qubit in optical tweezers.
Nat. Phys.
3
, 696699 (2007). doi:10.1038/nphys698
19. M. Schlosser, J. Kruse, C. Gierl, S. Teichmann, S. Tichelmann, G. Birkl, Fast
transport, atom sample splitting and single-atom qubit supply in two-
dimensional arrays of optical microtraps. New J. Phys.
14
, 123034 (2012).
doi:10.1088/1367-2630/14/12/123034
20. H. Kim, W. Lee, H.-g. Lee, H. Jo, Y. Song, J. Ahn, In situ single-atom array
synthesis using dynamic holographic optical tweezers. arXiv:1601.03833 (2016);
https://arxiv.org/pdf/1601.03833.pdf.
21. T. Grünzweig, A. Hilliard, M. McGovern, M. F. Andersen, Near-deterministic
preparation of a single atom in an optical microtrap. Nat. Phys.
6
, 951954
(2010).
doi:10.1038/nphys1778
22. B. J. Lester, N. Luick, A. M. Kaufman, C. M. Reynolds, C. A. Regal, Rapid
production of uniformly filled arrays of neutral atoms. Phys. Rev. Lett.
115
,
073003 (2015).
Medline doi:10.1103/PhysRevLett.115.073003
23. Y. H. Fung, M. F. Andersen, Efficient collisional blockade loading of a single atom
into a tight microtrap. New J. Phys.
17
, 073011 (2015). doi:10.1088/1367-
2630/17/7/073011
24. D. Jaksch, J. I. Cirac, P. Zoller, S. L. Rolston, R. Cote, M. D. Lukin, Fast quantum
gates for neutral atoms. Phys. Rev. Lett.
85
, 22082211 (2000). Medline
doi:10.1103/PhysRevLett.85.2208
25. M. Saffman, T. G. Walker, K. Mølmer, Quantum information with Rydberg atoms.
Rev. Mod. Phys.
82
, 23132363 (2010). doi:10.1103/RevModPhys.82.2313
26. H. Weimer, M. Müller, I. Lesanovsky, P. Zoller, H. P. Büchler, A Rydberg quantum
simulator. Nat. Phys.
6
, 382388 (2010). doi:10.1038/nphys1614
27. A. Browaeys, D. Barredo, T. Lahaye, Experimental investigations of dipole-dipole
interactions between a few Rydberg atoms. J. Phys. B
49
, 152001 (2016).
doi:10.1088/0953-4075/49/15/152001
28. J. D. Thompson, T. G. Tiecke, N. P. de Leon, J. Feist, A. V. Akimov, M. Gullans, A.
S. Zibrov, V. Vuletić, M. D. Lukin, Coupling a single trapped atom to a nanoscale
optical cavity. Science
340
, 12021205 (2013). Medline
doi:10.1126/science.1237125
29. A. Goban, C. L. Hung, S. P. Yu, J. D. Hood, J. A. Muniz, J. H. Lee, M. J. Martin, A. C.
McClung, K. S. Choi, D. E. Chang, O. Painter, H. J. Kimble, Atom-light interactions
in photonic crystals. Nat. Commun.
5
, 3808 (2014). Medline
doi:10.1038/ncomms4808
30. S. Murmann, F. Deuretzbacher, G. Zürn, J. Bjerlin, S. M. Reimann, L. Santos, T.
Lompe, S. Jochim, Antiferromagnetic Heisenberg spin chain of a few cold atoms
in a one-dimensional trap. Phys. Rev. Lett.
115
, 215301 (2015). Medline
doi:10.1103/PhysRevLett.115.215301
31. See supplementary materials on Science Online.
32. T. Xia, M. Lichtman, K. Maller, A. W. Carr, M. J. Piotrowicz, L. Isenhower, M.
Saffman, Randomized benchmarking of single-qubit gates in a 2D array of
neutral-atom qubits. Phys. Rev. Lett.
114
, 100503 (2015). Medline
doi:10.1103/PhysRevLett.114.100503
33. Y. Wang, X. Zhang, T. A. Corcovilos, A. Kumar, D. S. Weiss, Coherent addressing
of individual neutral atoms in a 3D optical lattice. Phys. Rev. Lett.
115
, 043003
(2015).
Medline doi:10.1103/PhysRevLett.115.043003
34. A. M. Kaufman, B. J. Lester, C. A. Regal, Cooling a single atom in an optical
tweezer to its quantum ground state. Phys. Rev. X
2
, 041014 (2012);
http://journals.aps.org/prx/abstract/10.1103/PhysRevX.2.041014.
35. J. D. Thompson, T. G. Tiecke, A. S. Zibrov, V. Vuletić, M. D. Lukin, Coherence and
Raman sideband cooling of a single atom in an optical tweezer. Phys. Rev. Lett.
110
, 133001 (2013). Medline doi:10.1103/PhysRevLett.110.133001
36. P. Kómár, T. Topcu, E. M. Kessler, A. Derevianko, V. Vuletić, J. Ye, M. D. Lukin,
Quantum network of atom clocks: A possible implementation with neutral atoms.
Phys. Rev. Lett.
117
, 060506 (2016). Medline
doi:10.1103/PhysRevLett.117.060506
37. J. F. Barry, D. J. McCarron, E. B. Norrgard, M. H. Steinecker, D. DeMille, Magneto-
optical trapping of a diatomic molecule. Nature
512
, 286289 (2014). Medline
doi:10.1038/nature13634
38. N. R. Hutzler, L. R. Liu, Y. Yu, K.-K. Ni, Eliminating light shifts in single-atom
optical traps.
arXiv:1605.09422 (2016); https://arxiv.org/abs/1605.09422.
39. D. Barredo, S. de Léséleuc, V. Lienhard, T. Lahaye, A. Browaeys, An atom-by-
atom assembler of defect-free arbitrary 2d atomic arrays. arXiv: 1607.03042;
https://arxiv.org/abs/1607.03042
40. C. Klempt, T. van Zoest, T. Henninger, O. Topic, E. Rasel, W. Ertmer, J. Arlt,
Ultraviolet light-induced atom desorption for large rubidium and potassium
magneto-optical traps. Phys. Rev. A
73
, 013410 (2006).
doi:10.1103/PhysRevA.73.013410
41. Y. R. P. Sortais, H. Marion, C. Tuchendler, A. M. Lance, M. Lamare, P. Fournet, C.
Armellin, R. Mercier, G. Messin, A. Browaeys, P. Grangier, Diffraction-limited
optics for single-atom manipulation. Phys. Rev. A
75
, 013406 (2007).
doi:10.1103/PhysRevA.75.013406
ACKNOWLEDGMENTS
We thank K.-K. Ni, N. Hutzler, A. Mazurenko, and A. Kaufman for insightful
discussion. This work was supported by NSF, CUA, NSSEFF, and Harvard
Quantum Optics Center. H.B. acknowledges support by a Rubicon Grant of the
Netherlands Organization for Scientific Research (NWO). During the completion
of this work, we became aware of a related approach (39).
SUPPLEMENTARY MATERIALS
www.sciencemag.org/cgi/content/full/science.aah3752/DC1
Materials and Methods
Figs. S1 to S5
Movies S1 to S3
References (40, 41)
17 June 2016; accepted 17 October 2016
Published online 3 November 2016
10.1126/science.aah3752
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 4
on June 7, 2018 http://science.sciencemag.org/Downloaded from
Fig. 1. Protocol for creating defect-free arrays. (A) A first image identifies
optical microtraps loaded with a single atom, and empty traps are turned off.
The loaded traps are moved to fill in the empty sites and a second image
verifies the success of the operation. (
B
) The trap array is produced by an
acousto-optic deflector (AOD) and imaged with a 1:1 telescope onto a 0.5 NA
microscope objective, which creates an array of tightly focused optical
tweezers in a vacuum chamber. An identical microscope objective is aligned
to image the same focal plane. A dichroic mirror allows us to view the trap
light on a charge-coupled-device camera (CCD) while simultaneously
detecting the atoms via fluorescence imaging on an electron-multiplied-CCD
camera (EMCCD). The rearrangement protocol is realized through fast
feedback onto the multi-tone radio-frequency (RF) field driving the AOD.
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 5
on June 7, 2018 http://science.sciencemag.org/Downloaded from
Fig. 2. Assembly of regular atom arrays.
(
A
) Single-shot, single-atom resolved fluorescence images
recorded with the EMCCD before (top) and after (bottom) rearrangement. Defects are identified and the
loaded traps are rearranged according to the protocol in Fig. 1, indicated by arrows for a few selected atoms.
(
B
) Two instances of successfully rearranged arrays (first two pictures), and one instance (last picture) where
a defect is visible after rearrangement. (
C
) The final arrangement of atoms is flexible, and we generate, e.g.,
clusters of two (top) or ten (bottom) atoms. Non-periodic arrangements and adjustable lattice spacings are
also possible. (
D
) High-resolution CCD image of trap array. Our default configuration for loading atoms
consists of an array of 100 tweezers with a spacing of 0.49 MHz between the RF-tones, corresponding to a
real-space distance of 2.6 μm between the focused beams (31).
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 6
on June 7, 2018 http://science.sciencemag.org/Downloaded from
Fig. 3. Quantifying the rearrangement performance. (A) The initial loading (blue circles) results in an
occupation probability of
0.6
for each trap in the array. After rearrangement (red circles), close to unity
filling is reached on the left side of the array. (
B
) In the initial image, the probability of finding a defect-free
length-N array (starting from the leftmost trap) falls off exponentially with N (blue circles). Following the
rearrangement of all loaded traps to form the largest possible array, we demonstrate strongly enhanced
success probabilities at producing defect-free arrays (red circles). Theory curves show limits set by the
total initial atom number (solid line), the background limited lifetime of τ = 6.2 s (dashed line) and the
product of both (dashed dotted line) (31). (
C
) Expected amount of time to wait on average to produce a
defect-free array of a given size taking into account the experimental cycle time of 200 ms (150 ms without
rearrangement). Without rearrangement, the wait time grows exponentially (blue circles). Employing the
rearrangement procedure, we can produce arrays of length 50 in less than 400 ms (red circles). All error
bars denote 68% confidence intervals, which are smaller than the marker size in (A) and (B).
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 7
on June 7, 2018 http://science.sciencemag.org/Downloaded from
Fig. 4. Creating and maintaining regular arrays using an atomic reservoir. (A) For a given target array
size, surplus atoms are kept in a reservoir and used for repetitive reloading of the array. (
B
) A 20 atom
target array with a reservoir of atoms on the right. Defects occasionally develop in the target array and are
replaced by atoms in the reservoir. The reservoir depletes as it is used to fill in defects. (
C
) By performing
repeated rearrangements (once every 50 ms) the probability to successfully produce a defect-free array
in any of these attempts increases and approaches the limit set by the number of initially loaded atoms
(dashed lines). We show data for targeting 40 (purple), 50 (yellow), and 60 (green) atom arrays. Solid
lines are guides to the eye. (
D
) Probing for defects and filling them once every 100 ms from the reservoir
extends the lifetime of a defect-free array. Shown is the success probability of maintaining arrays of 20
(circles) and 40 (triangles) atoms with (red) and without (blue) replenishing atoms from the reservoir.
With replenishing, the probability to maintain a defect-free array remains at a fixed plateau for as long as
we have surplus atoms in the reservoir. The initial plateau value is set by the probability that no atoms in
the array are lost in 100 ms (calculated value for 10 s single atom lifetime shown as the dotted line). All
error bars denote 68% confidence intervals, which are smaller than the marker size in (C).
First release: 3 November 2016 www.sciencemag.org (Page numbers not final at time of first release) 8
on June 7, 2018 http://science.sciencemag.org/Downloaded from
Atom-by-atom assembly of defect-free one-dimensional cold atom arrays
Vladan Vuletic, Markus Greiner and Mikhail D. Lukin
Manuel Endres, Hannes Bernien, Alexander Keesling, Harry Levine, Eric R. Anschuetz, Alexandre Krajenbrink, Crystal Senko,
published online November 3, 2016
ARTICLE TOOLS
http://science.sciencemag.org/content/early/2016/11/02/science.aah3752
MATERIALS
SUPPLEMENTARY
http://science.sciencemag.org/content/suppl/2016/11/04/science.aah3752.DC1
CONTENT
RELATED
http://science.sciencemag.org/content/sci/354/6315/1021.full
http://science.sciencemag.org/content/sci/354/6315/972.full
REFERENCES
http://science.sciencemag.org/content/early/2016/11/02/science.aah3752#BIBL
This article cites 37 articles, 5 of which you can access for free
PERMISSIONS
http://www.sciencemag.org/help/reprints-and-permissions
Terms of ServiceUse of this article is subject to the
registered trademark of AAAS.
is aScienceAmerican Association for the Advancement of Science. No claim to original U.S. Government Works. The title
Science, 1200 New York Avenue NW, Washington, DC 20005. 2017 © The Authors, some rights reserved; exclusive licensee
(print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement ofScience
on June 7, 2018 http://science.sciencemag.org/Downloaded from