24
Quantum Computing Experiments with Cold Trapped Ions

Ferdinand Schmidt‐Kaler and Ulrich Poschinger

QUANTUM, Johannes Gutenberg‐Universität Mainz, Institut für Physik, Staudingerweg 7, 55128 Mainz, Germany

24.1 Introduction to Trapped‐Ion Quantum Computing

24.1.1 History of Single Ion Trapping

The idea for dynamic trapping in alternating fields was conceived by Paul et al. in 1953 (1) and rewarded with the Nobel prize in 1989 (2). As static electric fields do not allow for trapping of charged particles, his invention was to employ an oscillating quadrupole field, which can result in bounded and stable trajectories. This realized a possibility to confine charged particles in deep potential at long trapping times, which has led to experimental progress in many fields of physics. While Paul was probably inspired by the alternating focusing and defocusing elements typically used in storage rings for high‐energy and nuclear‐physics research, his invention finally initiated a rapid development in ultra‐cold atomic and molecular quantum physics. A further driving force of this field was the development of tunable laser sources, which led to the invention of laser cooling for neutral atoms by Hänsch and Schawlow (3), and at the same time for trapped ions by Wineland and Dehmelt (4).

Experiments with single trapped and laser‐cooled ions started as early as 1980, when the fluorescence of one single laser‐cooled Baimages ion was observed (5). At that time, research was motivated by two goals: To beat the existing limitations in the spectroscopic precision of atomic clock transitions, and to manipulate and observe a single atomic system in well‐controlled interaction with an optical laser field – Gedankenexperimente of quantum optics became experimental reality. This triggered a stimulating contact between theory and experiments in quantum optics and atomic physics. Highlights of that era include the demonstration of quantum jumps with a single ion (6,7), the investigation of photon antibunching in emitted resonance fluorescence light (8), and the demonstration of the coherent dynamics of a driven two‐level system, realizing the famous Jaynes–Cummings model. In this model of a two‐level atomic system coupled with the equidistant ladder of states of a harmonic oscillator, the interaction of both systems leads to a periodic exchange of excitation, known as Rabi oscillations. The spin‐1/2 system in nuclear magnetic resonance experiments by Felix Bloch, Isidor Rabi, and Edward Hahn is equivalent to an atomic two‐level system. But, only very few systems at that time reached the conditions of “strong coupling,” where the coherent interaction strength exceeds the dissipative rates such as spontaneous atomic decay: Only in 1990, experiments with Rydberg atoms interacting with superconducting cavities that support a set of quantized electromagnetic field modes (911) could reach this regime, at the same time, the equivalent coherent Jaynes–Cummings dynamics was demonstrated with a laser‐driven single ion trapped in a harmonic potential resulting in quantized eigenmodes of vibration (12). Laser cooling of a single trapped ion into the vibrational ground state of motion was demonstrated (13), and this degree of control for both the motional degree of freedom and the two‐level system paved the way for seminal quantum experiments: Using a single trapped ion, the NIST, Boulder group led by David Wineland was able to create a Schrödinger cat state of a trapped ion's oscillatory motion (14). Serge Haroche and his team were able to prepare a Schrödinger cat state of the photon field of a cavity and observe its successive decay – decoherence due to coupling to the environment (15). The 2012 Noble prize for physics was awarded to both of them “for ground‐breaking experimental methods that enable measuring and manipulation of individual quantum systems” (16). This groundbreaking success stimulated both theoretical concepts and proposals as well as experimental progress for quantum computing with trapped ions.

24.1.2 History of Quantum Computing

The very idea of quantum computing, when pioneered by Manin (17,18), Feynman (19,20), and later by Deutsch (21) in the 1980s, was known to only a small number of insiders. The interest in quantum computing strongly increased when Peter Shore invented the factorization algorithm (22), as the important implications for modern data encryption systems were obvious. In 1994, at the occasion of the 14th International Conference on Atomic Physics, Ekert (23) brought this factorization algorithm (24) into discussion. A quantum bit, or qubit, allows for encoding superposition of information in an atomic two‐level system. Logic gate operations (quantum gates) control the state of single or multiple qubits. Such operations can generate multiqubit entanglement. The need for a clear concept of an experimental platform based on quantum optics fostered many theoretical and experimental research activities. And precisely, this first “blueprint” of a future quantum computer was given by Ignacio Cirac and Peter Zoller in their seminal proposal for a quantum‐logic two‐qubit gate in 1995: Each ion in a linear crystal of images ions stores one bit of quantum information in two long‐lived electronic levels, referred to as images and images . Quantum gates are implemented by laser‐ion interactions (25). The logic state of such a qubit can thus be expressed as a general superposition images with complex amplitudes images for which images holds. Experimentally, a first quantum logic operation on a single ion was shown (26), and in 2003, the proposed CNOT gate operation was demonstrated with a pair of ions that represent the control‐ and the target‐qubit (27). The Cirac–Zoller proposal stimulated a series of experiments using trapped ions for quantum information processing. We can claim today that the principle of a QC is proven and that trapped ions are indeed a pioneering experimental platform for its development. This experimental development also helped to prevail against an initially overwhelming criticism concerning any experimental realization of a quantum computer (28).

24.1.3 Recent Milestones in Ion Trap Quantum Computing

Since the aforementioned pioneering experiment, the field has seen rapid progress, with many new proposals and demonstrations for quantum gate operations. Highlights by the Ion storage group at NIST Boulder and the team at Innsbruck University led by Rainer Blatt have been the unconditional teleportation with massive and long‐lived carriers of qubits (2931), Bell tests to support quantum theory against local‐realistic theories (32) and multiparticle entanglement with up to 14 qubits (33). Furthermore, novel types of quantum gate operations, either with microwaves (3437) or with short laser pulses (38) have been shown. Recently, the Shor factoring algorithm was realized (39), and a topologically protected qubit (40) based on seven physical qubits (ions) has been demonstrated. A set of methods for scalable ion trap quantum information processing was demonstrated (41). Today, two‐qubit entangling gate fidelities of 0.999 ± 0.001 (42,43) are realized, such that quantum error correction beyond proof‐of‐principle demonstrations seems to be within reach.

Trapped ions feature several advantages which make them one of the leading approaches for a realization of a future quantum computer (QC):

  • In an ion trap, before a quantum algorithm starts, the quantum state of each ion in the crystal can be prepared such that the register of qubits, a quantum register, is initialized. The necessary techniques for this are laser cooling and optical pumping. Both techniques have been established already for the purpose of high‐resolution laser spectroscopy. More recently, sympathetic cooling methods have been developed, such that the motional state of a quantum register may be cooled without affecting its coherence.
  • Ions can be stored in a linear Paul trap, or a segmented multizone trapping device, which use the charge of the ions as a “handle” via interaction with electric fields. Therefore, the trapping potential is very deep and tight, without cross‐talk to the internal electronic states – which are employed to store qubit information. Cirac and Zoller deduced that in an experiment with a trapped ion crystal serving as a quantum register, with the ions being held in a Paul trap under ultra‐high vacuum conditions, one could avoid undesired coupling of this quantum register to the environment. Indeed, measured qubit coherence times exceed the time for single qubit operations by more than five orders of magnitude.
  • For the necessary quantum gate operations, ions are coherently manipulated by laser radiation. For two‐qubit gates, the coupling to a common vibrational mode – termed quantum bus – is employed. Various two‐qubit gate operations are proposed and a number of them have been realized experimentally (see Section 24.4.5).
  • Fairly unique for trapped ions is the asset of a very high readout efficiency and fidelity. The technique of “electron shelving” dates back to Dehmelt (44). It allows to scatter a sufficiently large number of photons off an ion and to observe at least some fraction of them as clicks on a detector, if the ion is in qubit state images , while no photons are scattered if the ion is in state images . A single‐shot readout fidelity of better than 99.93images has been reached (45).

Today, the central challenges in trapped ion quantum computing are (i) developing architectures for scalable quantum computing with trapped ions and (ii) optimizing the gate fidelity and speed, and (iii) implementing quantum error correction schemes such that a large number of operations are feasible.

This article is organized as follows: After a discussion of linear Paul traps and quantized eigenmodes of ion crystals in the harmonic trap potential, we explain the operations on a static linear qubit registers. Here, tightly focused laser beams allow for single qubit addressing. Several highlights of ion trap quantum computing have been realized with this architecture. Modern segmented trap devices are aiming for scalable quantum computing toward a much larger number of trapped ions in reconfigurable quantum registers. This architecture has been coined quantum charged coupled device ( CCD ) (46). In the following, we discuss ion–laser and ion–microwave interactions and give examples for ion species used for qubits. We outline single‐qubit gate operations and a number of different two‐qubit gate schemes. The experimental realization of the Cirac–Zoller gate is described. Quantum logic operations have been combined in different ways to establish quantum algorithms. In order to exemplify the realization of elementary quantum algorithms, we focus on quantum teleportation. The article will sketch the most recent highlights and discuss the future challenges.

24.2 Paul Traps

Charged particles, such as atomic ions, can be confined by electromagnetic fields, either by using a combination of static electric and magnetic fields (Penning trap) or by a time‐dependent inhomogeneous electric field (Paul trap) ( 1,47). In the latter case, an ac‐electric field is generated by an appropriate electrode structure and creates a ponderomotive pseudo‐potential which can confine charged particles. The motion of a particle confined in such a field involves a fast component synchronous to the applied driving frequency (micro motion) and the slow harmonic (secular) motion in the dynamical pseudo‐potential.

In order to confine particles in a harmonic potential, we require a restoring force which increases linearly with the distance from the origin of the trap. Such an effective force is generated by a quadrupole potential images , where images denotes a voltage applied to a quadrupole electrode configuration, images is the characteristic trap size and the constants images determine the shape of the electric potential images , given by the solution of Laplace's equation images . For example, in the case of a three‐dimensional electric field, we find images . The potential is confining in the images and images directions, but anticonfining along the images direction. Therefore, a static electric field alone cannot lead to three‐dimensional confinement. If, however, an alternating electric field is applied, the resulting potential is attractive in the images and images directions for the first half cycle of the field, and attractive in the images direction for the second half cycle. A suitably chosen amplitude and frequency images of this alternating field then allows for trapping of charged particles of mass images and charge images , in all three dimensions. The three‐dimensional Paul trap provides a confining force with respect to a single point in space, the node of the rf‐field, and therefore is mostly used for single ion experiments or for the confinement of three‐dimensional crystallized ion structures.

If we consider the electric rf‐field in a two‐dimensional geometry along images and images axis only, we find images . Now, the potential is attractive in images and repulsive in the images direction in the first half cycle of the ac‐field, and vice versa for the second half. This property is well known from the quadrupole mass filter. Here, confinement of a charged particle is given only in the (radial) images and images directions. If an additional static dc‐potential is applied in images direction, the particle is trapped radially and axially, and we may talk about a linear ion trap. In order to realize a quantum register with trapped ions, a linear arrangement of the ions (i.e., ion strings) is advantageous. This geometry allows for individual observation of the ions and individual coherent manipulation of an ion's quantum state.

24.2.1 Stability Diagram of Dynamic Trapping

We focus first on the case of a two‐dimensional trap and discuss the parameter range for dynamic trapping. To confine the ions in 2D, we apply an rf‐voltage images and an (optional) dc‐voltage images to the trap electrodes. Near the trap axis for images , this gives rise to an alternating electrostatic potential of the form

24.1 equation

where images denotes the distance between the trap axis and the surface of one of the electrodes. The equations of motion in dimensionless form resulting from 24.1 are the Mathieu equations,

24.2 equation
24.3 equation

where images and images with images . The general solution of Eqs. 24.2 and 24.3 can be given as an infinite series of harmonics of the trap frequency images (47). For the appropriate choice of parameters images and images , the ion trajectory is bounded in space and momentum: dynamical trapping is achieved, see Figure 24.1. If the conditions images hold, an analytical approximate solution to the equations of motion can be given. It consists of a harmonic secular motion (macromotion) at frequencies images with a superimposed micromotion at the trap drive frequency images ,

24.4 equation

The amplitude images and the phases images depend on the initial conditions, and the secular frequencies are given by

24.5 equation
Illustration of Stability diagram for a linear quadrupole configuration.

Figure 24.1 Stability diagram for a linear quadrupole configuration. Inside the red boundary lines, the ion trajectory is confined within the trap volume. A similar stability diagram exists for the images ‐direction. Inset: Ion trajectory as solution of the Mathieu equation, plotted versus time with images  = 0 and images  = 0.2. The oscillation exhibits a slow secular motion at frequency images , superimposed with the fast micromotion at the frequency images of the rf‐drive.

24.2.2 3D Confinement in a Linear Paul Trap

Axial confinement is provided by an additional static potential images applied along the images ‐axis using additional axial electrodes. This gives rise to an electrostatic harmonic potential well along the images ‐direction, which is characterized by the longitudinal trap frequency

24.6 equation

Here, z0 is half the length between the axial confinement electrodes, and images is a factor of order unity which accounts for the specific electrode geometry. Values of images can be obtained either numerically or, in some cases, analytically. For the macroscopic ion trap in Figure 24.2a), a voltage of images  = 2000 V applied to the tips gives rise to an axial trap frequency of images 1.4 MHz for images . Under typical operating conditions, radial trap frequencies of about images 5 MHz are achieved. Until today, groundbreaking QC demonstration experiments with linear static qubit registers have been carried out by the Innsbruck group in this device.

Illustration of Different realizations of Paul traps.

Figure 24.2 Different realizations of Paul traps. Each panel contains a sketch of the fundamental electrode geometry of the respective trap type. (a)A macroscopic linear endcap Paul trap from, for example, (48), where linear ion crystals are stored in a single trap potential. (b) Three‐dimensional segmented microchip linear traps, where the electrostatic potential along the trap axis is controlled by the individually applied voltages on the dc segments. Ions can be simultaneously stored at different trap sites and shuttled within the trap by changing the dc segment voltages. (c) Surface electrode traps. The sketch indicates equipotential lines of the rf electrodes, leading to an alternating quadrupole above the surface. Modern fabrication methods allow for complex electrode geometries and features such as slits, loading holes, and junctions with optimized transfer geometry. Cutting edge traps feature signal routing within the trap structure, which allows for contacting island electrodes.

The resulting (pseudo)potentials in all spatial directions are harmonic, and the motion of a trapped ion is accurately described by a quantum harmonic oscillator with frequencies images . It is a major advantage of linear traps that the radial and axial trap frequencies can be adjusted freely and independently by tuning the applied voltages. Further details of the calculation of the stability diagram for 3D linear traps are given in Ref. (49).

24.3 Ion Crystals and Normal Modes

The first two‐qubit quantum gate operations were demonstrated in a linear trap. Here, ions can be confined and optically cooled such that they form ordered structures (5053) with a fixed equilibrium position of each ion – termed Coulomb crystals. If the radial confinement along the images ‐ and images ‐axis is strong as compared to the axial confinement, that is, images , ions arrange in a linear crystal along the trap images ‐axis at distances determined by the equilibrium of the Coulomb repulsion and the harmonic potential providing axial confinement. An example of a string of ions in a linear Paul trap is shown in Figure 24.3. The average distance between two ions, in this case, is about 10 images m. The importance of normal modes in the ion‐trap quantum computer is based on the fact that all two‐qubit gate operations rely on the excitation of common vibrational degrees of freedom of the linear ion crystal using the motion as a quantum bus.

Illustration of Linear crystal with eight 40Ca+ ions imaged by a CCD system.

Figure 24.3 Linear crystal with eight images Caimages ions imaged by a CCD system. The fluorescence light near 397 nm is observed while the ions are excited by laser light at 397 and 866 nm. The trap used for this is shown in Figure 24.2a). The exposure time for the CCD image of the ion fluorescence was 1 s, the resolution measured for the imaging system consisting of lens and CCD camera was better than 4 images m.

Consider images ions in a linear arrangement, where the position of the images th ion is denoted by images , see Figure 24.4. The ions experience the trap potential and their mutual Coulomb repulsion. The total potential energy is given by

24.7 equation

The first term describes the potential energy in the harmonic trap, while the second describes the mutual Coulomb repulsion of the ions. For simplicity, both radial frequencies are assumed to be equal, that is, images , where images is a parameter that describes the anisotropy of the trap. The equilibrium positions images of the ions in a crystal follow from the condition

Image described by caption and surrounding text.

Figure 24.4 Illustration of the model for the linear crystal, position vector components images for the images th ion, and its equilibrium position (images ) as indicated by a cross (X).

24.8 equation

with images . The values of equilibrium positions can be determined numerically (54,55). It is convenient to introduce a dimensionless length scale

24.9 equation

An analytic approximation for the minimum interion distance in a string of images ions yields images . At an axial frequency of images =700 kHz, the distance between two images Caimages ions is 7.6 images m and reduces to 6 images m in case of three. Small deviations of the ions from their equilibrium positions are described by images , and we will see that the motion can then be described in terms of normal modes of the entire chain oscillating at distinct frequencies ( 55,56). The potential energy of the Coulomb crystal is now written as:

24.10 equation

A Taylor expansion up to second order in the deviations images around the equilibrium positions is employed to obtain an effectively harmonic Coulomb interaction (57):

24.11 equation

The first line of Eq. 24.11 describes the potential along the (axial) images ‐direction only, while the second line describes the potentials along the (radial) images ‐ and images ‐directions (58).

In the linearized model, the eigenmode frequencies and eigenvectors are found by diagonalization of Hessian matrices images for axial and images for the radial directions:

24.12 equation
24.13 equation

with the dimensionless equilibrium positions along the axial direction images . The eigenfrequency of the images th normal mode along direction images is given by images , where images is the center‐of‐mass mode frequency along direction images , and images is the images th eigenvalue of the respective Hessian. The physical meaning of the eigenvectors is as follows: The images th component of the images th eigenvector along images indicates the direction and amplitude of oscillation of the images th ion at excitation of the images th collective mode along images , with respect to the other ions. As an example, for the center‐of‐mass mode, all components of the corresponding eigenvector have the same modulus and sign, therefore all ions oscillate identically. The eigenfrequencies of linear crystals comprised of images ions are shown in Figure 24.5. For an increasing number of ions, the frequency differences become smaller. Therefore, the selected quantum bus vibrational mode becomes less well separated in frequency from the other modes, which represents a problem of scalability of the Cirac–Zoller 1995 quantum gate proposal (Section 24.4.5.1). This problem does not occur for the approach where a reconfigurable quantum register is employed. Here, only a small number of ions are exposed to the laser that drives the quantum gate operation. During gate operations, all other ions which do not participate are stored in distinct potential wells.

In the architecture with static quantum registers, one of the axial normal modes is used for the quantum bus, for example, the center‐of‐mass oscillation at images corresponding to an oscillation of the entire chain of ions moving back and forth as if they were rigidly joined. The second normal mode corresponds to an oscillation where the ions move in opposite directions. More generally, this so‐called breathing mode describes a string of images ions with each ion oscillating at an amplitude proportional to its equilibrium distance from the trap center. The vibrational modes are quantized in the familiar way by introducing operators for momentum and position, together with the canonical commutation relations (55).

Illustration of Eigenfrequencies of an N = 3, 5, and 10 ion crystal.

Figure 24.5 Eigenfrequencies of an images ion crystal. Black, radial modes; gray, axial modes. The axial trap frequency is images  = 1 MHz and the radial trap frequency is images  = 5 MHz.

We summarize the most important results of the explicit calculation ( 55 57) of the axial normal modes linear ion crystals consisting of images ions: (i) Exactly images axial normal modes and normal frequencies exist. (ii) The center‐of‐mass mode frequency is the lowest frequency, and it is equal to the frequency of a single ion. (iii) Higher order axial frequencies are almost independent of the ion number images , and are given by images (1, 1.732, 2.4, 3.05(2), 3.67(2), 4.28(2), 4.88(2), …), where the numbers in brackets indicate the maximum frequency deviation as images increases from 1 to 10 ions. (iv) Even though the Coulomb interaction results in a non‐linear force, the ions undergo harmonic oscillations about their equilibrium positions. And even though the ion trajectories in dynamical Paul trapping potential are described by Floquet equations, the harmonic oscillator approximation holds astonishingly well for most of the situations (57).

24.4 Trap Technology

In order to realize a quantum computing device which clearly exceeds the capabilities of classical computers, the fundamental question lies in the scaling to large amounts of qubits. For trapped ions, conventional linear Paul traps do not offer the prospect of scalability for several reasons: First, we have seen in Section 24.3 that the confinement of ion strings of increasing length leads to spatial and spectral crowding, which in turn leads to the deterioration of quantum gate fidelities. Furthermore, longer strings lead to trap instabilities, and the addressing and readout of qubit ions become impracticable beyond small register sizes.

24.4.1 Trap Architectures

The scalability perspective for trapped ions was opened up with the seminal proposal for a quantum CCD (46), where – similar to a CCD in modern cameras – ions are moved within a distributed architecture by changing control voltages on electrodes. The underlying devices are segmented ion traps, that is, devices which consist of a multitude of trapping electrodes, arranged in a geometry which allows for trapping ions at different locations and for shuttling the ions between these zones. In the last decade, such traps were developed and successfully demonstrated by several research groups, and the technology and required methods have reached a maturity which already allows for conducting state‐of‐the‐art experiments and quantum algorithms with few‐qubit systems. As segmented ion traps are produced using micro‐fabrication techniques, this also offers the possibility of miniaturization, which is a natural requirement for scalability. A common challenge for miniaturized trap lies in anomalous heating: Due to the increased proximity of the ion to surfaces, electric field noise generated by these leads to undesired ion motion leading to thermalization with the environment. The experimentally heating rates lie orders above the limit given by Josephson noise (59), which leads to the conclusion that the noise is generated by surface contaminants or structural defects within the metal layers. As pointed out in Section 24.4.5 , all entangling gate schemes known so far rely on control over the ion motion in the quantum regime. Therefore, significant research efforts are devoted to understanding and mitigating these effects, for example, by using ion traps in cryogenic environments (6062) or in‐situ surface cleaning via ion bombardment (63,64).

Currently, segmented microstructured ion trap fall into two different categories.

24.4.1.1 Multilayer Sandwich Traps

These traps consist of a stack of wafers, into which a trap slit is crafted. The wafers are metalized such that distinct control electrodes are constituted, which can be individually connected to external control electronics. The wafers have to be aligned and fixed, such that an electrode geometry rather similar to a linear Paul trap is obtained. The fabrication of the structured wafers is accomplished via laser cutting, where different technological approaches can be employed. The resulting structure sizes are typically limited to more than 10 images m. The metalization of the wafers is accomplished via evaporative coating. A widely used trap material is gold, which requires an additional adhesion layer of, for example, titanium to stick on typical wafer material such as alumina, aluminum nitride, glass or quartz. Such coatings achieve metal layers less than a micrometer thickness. Optionally, a thick metal film can be deposited by subsequent electroplating, for which the surface layers need to be electrically contacted.

This type of trap has the advantage that the trap potentials closely resemble these of the original Paul trap, thus deep and strong confinement at electrode–ion distances in the 100–500 images m range is possible. Trap frequencies of several megahertz are achieved for medium‐weight ion species and for trap voltages far from electric breakdown risk. However, the resemblance to conventional Paul traps also limits the options for scalability: While these traps can serve to establish fundamental working principles of the quantum CCD, the complexity of the possible trap structures is limited by the complexity of fabrication. Several research groups work on adopting fabrication methods for semiconductor microstructures for enabling monolithic three‐dimensional segmented trap at increased structural complexity and reduced dimensions (6567).

24.4.1.2 Surface Electrode Traps

Here, all electrodes are arranged in a two‐dimensional plane. For suitable geometry parameters, a quadrupole potential exists above the surface, yielding a pseudo‐potential minimum for stable trapping. Such traps are fabricated using lithographic techniques adapted from semiconductor fabrication processes. This allows for almost arbitrarily complex structures, including segmentation, junctions, and geometries varying across the trap structures according to the purpose of the respective trap region. In this respect, surface electrode traps offer much better prospects for scalability. Since the demonstration of the first surface electrode trap by the NIST group (68), these traps have evolved toward complex structures featuring several junctions and up to 150 controllable trap zones (69). In modern traps, slits along the trap axis allow for better optical access, elevated trap electrodes allow for shielding isolating trenches, and routing layers below the electrode plane allow for connection and control of island electrodes.

However, these traps are more complicated to operate than their three‐dimensional microtraps: The pseudo‐potential minimum above the surface is way more shallow and asymmetric, which increases trap loss rates and decreases trap frequencies. In order to achieve sufficiently tight trapping conditions, the ions have to be trapped rather close to the surface, often at distances of below 100 images m. In order to suppress anomalous heating due to electric field noise generated by the trap surface (59), and to minimize trap losses from background gas collisions, such ion traps are often used at cryogenic temperatures.

24.4.2 Ion Shuttling

Ion shuttling operations in segmented trap are performed in essentially the same way as photo‐induced charges are transferred for readout on a CCD chip: By sweeping the control voltages applied to neighboring trap electrodes such that the resulting confining electrostatic potential well is moved, the confined ions are moved as well. In the frictionless case of trapped ions, this is physically equivalent to moving the support of a pendulum. The challenge behind these operations lies in the requirement to perform these operations fast – on the timescale set by the trap frequencies – in order to not compromise the quantum computer operation by excessive overhead. However, the excitation of ion motion which persists after the shuttling is to be avoided, as it would deteriorate the fidelity of subsequent gates. Thus, fast shuttling operations pose stringent requirements on control, especially in terms of the signal integrity of the utilized voltage ramps. To that end, signal generators have been developed which provide arbitrary waveform generation for a large number of up to 60 independently controllable output channels, which are updated in real time at rates which exceed typical trap frequencies, that is, more than 10images samples per second. At the same time, these signal generators feature excellent noise characteristics permitting the desired degree of control (70,71).

Following initial demonstrations of shuttling operations (72), diabatic shuttling at duration of a few trap periods has been accomplished at residual motional excitations below the single‐quantum level (73,74). Furthermore, more complex movement operations such as separation and merging of ion crystals have been shown ( 73,75). These operations are significantly more challenging, as they involve the transition between a common single‐well potential and a double‐well potential. This leads through a situation with low‐confinement along the movement axis. Precise calibration and optimized voltage waveforms are required (76) to avoid strong excitation either from increased thermalization rates at low trap frequencies or from oscillatory excitation due to insufficient control over the process.

24.4.3 Ion–Light Interaction

As the relevant theory is outlined in Section 23 and in review papers (77,78), we concentrate on examples of carrier and sideband Rabi oscillations to demonstrate the Jaynes–Cummings dynamics. The coupling between two electronic states images and images is mediated by a light field, characterized by the Rabi frequency images and resulting in the Hamiltonian images . If the laser frequency fulfills the resonance condition for the bare electronic transition images , the interaction is not affecting the vibrational modes (carrier transition). Vibrational modes can also be coherently excited by the laser–ion interaction. The exchange of momentum between the ion crystal and the optical field is governed by the Lamb‐Dicke factor. For a single ion and coupling to a single vibrational mode, this is given by ratio of the ion's position‐space ground‐state wavefunction and the wavelength images of the driving radiation:

24.14 equation

where images is the ion mass, images is the eigenfrequency of the respective normal mode, and images is the effective wavenumber of the optical field, where an angle between images between the direction of vibration of the normal mode and the propagation direction of the light is taken into account.

Including coherent coupling to the motion, the carrier Rabi frequency images remains almost unchanged for small values of images . However, if the laser excitation frequency is tuned to images , that is, it is red detuned with respect to the carrier frequency by trap frequency, images , we realize the Hamiltonian of the Jaynes–Cummings type. For the ion initially prepared in the Fock state images of the harmonic motion, the red sideband Rabi frequency is given by to images . Blue laser detuning by the trap frequency, with images , realizes the anti‐Jaynes–Cummings Hamiltonian with images . For resolving the carrier and sideband transitions spectroscopically, we require an optical transition between long lived electronic states images and images with a linewidth much smaller than the trap frequency.

Illustration of Coherent dynamics on the S1/2 to D1/2 optical qubit transition in 40Ca+.

Figure 24.6 Coherent dynamics on the images to images optical qubit transition in images Caimages , see Figure 24.7. We identify the images state with images and the Dimages with images . (a) Single ion Rabi oscillation on the carrier transition. Starting from images at images , the state images is reached near images 0.5 images s. After a images  = 2images rotation near images 0.9 images s, the state images is reached, and only after a full images rotation for images s, we recover the initial quantum state images . Here, a Rabi frequency of images 1.09 MHz is reached. (b) Rabi oscillation on the blue sideband at laser detuning from the carrier transition of images . The initial quantum state images evolves into as superposition images after an interaction time of about 40 images s. The state images is reached for a images ‐pulse at about images 75 images s. Here, the sideband was driven with images 7.4 kHz. With respect to the carrier transition, the Rabi frequency of the sideband dynamics is reduced by the Lamb‐Dicke parameter images , which is about 4% for the data shown here.

In experiments, one particular vibrational mode is selected as the quantum bus mode. Criteria for this selection are a low motional heating rate, that all ions to be addressed participate in the collective vibration, and sufficient spectral separation from other modes. For experiments that demonstrate the interactions for quantum gates, a single ion is kept in a Paul trap and the center‐of‐mass mode at images is used. An experimental four‐step‐sequence is applied:

  1. The ion is laser cooled into, or near to the ground state of vibration images .
  2. The electronic state is initialized to the ground state images by optical pumping.
  3. Laser light is applied with a fixed laser frequency, phase and intensity and interaction time images . Thus, a superposition of electronic qubit basis states images is prepared.
  4. The ion is exposed to light resonant with a dipole transition. State‐dependent resonance fluorescence is observed only if the ion is projected into images . If light is detected, the ion is measured to have been in state images , while the detection of no fluorescence means the ion is measured to have been in state images .

Finally, the probability for upper state population images is revealed as the average of all measurement outcomes in the above sequence (a)–(d) for a large number of repetitions.

Figure 24.6a exemplifies single qubit (carrier) rotations denoted by images , see Section 23, where the pulse area images is the product of pulse duration and Rabi frequency, and images is controlled by the phase of the laser field. Rotations on the blue sideband Rabi are denoted by images , and this operation is used to coherently transfer quantum information from the electronic degree of freedom onto the quantum bus vibrational mode. In the following section, we discuss possible qubit candidates and the corresponding experimental realizations.

Illustration of 40Ca+ (a, b) and 9Be+ (c) level schemes.

Figure 24.7 images Caimages (a, b) and images Beimages (c) level schemes, as examples for three types of qubits. The wavelengths of the different transitions are indicated. (a) For images Caimages , the lifetimes of the ion in the excited state Dimages is images 1.2 s. A laser near 729 nm serves to drive the optical qubit transition. (b) The spin–qubit employs the sublevels ground states, that is, the spin of the valence electron, to store quantum information. It is manipulated via a stimulated Raman transition near 397 nm, far off‐resonant to the Simages Pimages dipole transition. (c) In images Beimages , the two hyperfine qubit levels are employed, and coherent manipulation is carried out either via a stimulated Raman transition near 313 nm, or via a direct microwave transition between both hyperfine states.

24.4.4 Levels and Transitions for Typical Qubit Candidates

Although an ion trap is deep and capable of holding every kind of atomic ion, only a few atomic species are actually suitable for QC experiments. These ions should exhibit energy levels appropriate for the implementation of a stable two‐level system with long‐lived qubit levels images and images , and the ion has also need to have a closed transition to a short‐lived excited state to allow for laser cooling and efficient fluorescence detection. The “ideal ion” typically has one electron in the outermost shell (hydrogen‐like electronic structure) and a correspondingly simple electronic level structure. The two‐level qubit system can either be provided by two hyperfine ground states, by Zeeman sublevels or by a long‐lived metastable electronic state. Prominent examples are the hyperfine‐qubit in images Beimages (pioneered by the NIST, Boulder group), the so‐called optical qubit in images Caimages , where the ground state Simages and the optically excited metastable level Dimages are used (pioneered by the Innsbruck group) or the spin‐qubit, where information is stored in the two Zeeman sublevels of the images Caimages Simages ground state (pioneered by the Oxford group and currently used by the Mainz group). The level schemes of images Caimages and images Beimages are shown in Figure 24.7. Widely employed qubit implementations use the hyperfine ground states of images Ybimages or images Mgimages , and the isotope images Caimages (36,79). Qubit manipulations via stimulated Raman transitions feature the advantage that both required light fields can be derived from one single laser source, such that the differential phase fluctuations of both beams can be kept very small at moderate experimental effort. If the detuning of these beams from the resonance is chosen large enough, spontaneous emission from the off‐resonantly excited Pimages and Pimages states is suppressed and the coherence of the qubits is hardly affected (80). Additionally, the hyperfine and the spin‐qubits work with electronic ground states that do not show spontaneous decay. For this case, the decoherence is dominated by magnetic field fluctuations, but using magnetic shielding (81) or magnetic‐field insensitive clock states, the coherence times can exceed a few seconds. Microwave qubit operations on hyperfine qubits have been demonstrated as an alternative to optical qubit manipulations in images Ybimages , images Caimages , and images Beimages ( 34 37).

24.4.5 Multiqubit Entangling Gates

In this section, we explain entangling gate schemes which have been implemented at high fidelity. We first describe the Cirac–Zoller gate scheme because it illustrates how laser‐ion interaction can be employed to generate entanglement.

24.4.5.1 The Cirac–Zoller Scheme

An entangling quantum gate between the internal states images of any pair of ions images and images in a linear string can be achieved by three successive laser‐driven operations, addressing the images th, then the images th, and finally again the images th ion (82). The gate operation relies on the initialization of the ion crystal in the vibrational ground state of the quantum bus mode images ( 26 8385) of the quantum bus mode and individual optical addressing of ions ( 27,86), which are technically demanding requirements. However, the gate operation can be easily understood by looking at the stepwise flow of quantum information:

  1. First, the quantum state of the control ion is mapped onto the bus mode by a images pulse on the blue sideband of the qubit transition,
  2. A controlled‐NOT gate is carried out between the bus mode the target ion, driven by a images rotation on an auxiliary transition, which is conditional on the vibrational quantum state (see Figure 24.8a))
  3. The state of the bus mode is mapped back onto the control ion by a second images pulse on the blue sideband, such that the final motional state is disentangled with the state of the qubits and the state of the control qubit remains unaffected.
Image described by caption and surrounding text.

Figure 24.8 (a) Level scheme for the Cirac–Zoller gate scheme: A red sideband 2images ‐pulse on the auxiliary transition is driven, such that the state images acquires a phase factor of images 1. Ellipses indicate Rabi cycles (dashed: no resonant level is available and thus no phase accumulation takes place, solid: 2images phase accumulation). (b) Composite phase gate (87). A blue sideband 2images ‐pulse is driven, and all states acquire a phase factor of images 1 – except for the state images , which is decoupled from the blue sideband. The dashed box indicates the set of computational basis states.

24.4.5.2 Experimental Realization of the Cirac–Zoller Gate

The operation of mapping quantum information between the control qubit and the bus mode are simple images pulses. We therefore focus on the central controlled NOT operation between the bus mode and the target ion. This operation is further decomposed into two Ramsey images rotations on the carrier transition of the target qubit, which map a phase accumulated between the two pulse onto resulting population. Between the Ramsey pulses, a conditional phase is accumulated in the course of a controlled‐phase gate. Defining the computational subspace by images , this controlled‐phase gate can be described by a diagonal unitary evolution matrix images . Unlike the original proposal (82), which requires a images ‐rotation on an auxiliary transition, the experiment ( 27,88) uses a blue sideband excitation leading to pairwise coupling between the states images except for the state images , see Figure 24.8b). In this case, the evolution of the controlled phase gate in the relevant subspace reads images . For every basis state for which the controlled‐phase gates yields a resulting phase factor of images , the state of the target qubit is not flipped with respect to its original state after the second Ramsey pulse.

Illustration of state evolution for the composite phase gate, Rphase visualized on the Bloch sphere.

Figure 24.9 The state evolution for the composite phase gate, images is visualized on the Bloch sphere. (a) Bloch sphere for the two‐level system images . The initial state is images , indicated by the black arrow. Pulse images rotates the state vector about the images ‐axis by images . images accomplishes a images ‐rotation about the images ‐axis. It therefore transforms the state to its mirror image about the images images ‐plane. Consequently, images rotates the state vector all the way down to the bottom of the sphere. images represents a images ‐rotation about the images ‐axis. The final state is identical to the initial one, except the acquired phase factor of images . (b) The same laser pulse sequence acting in the images subspace. Again, the final state is identical to the initial one, except for the acquired phase factor of images .

This means that for the controlled phase gate, we perform an effective images pulse (48) on the two two‐level systems (images and images , which flips the sign of the state for all computational basis states except for images ). Since the Rabi frequency depends on images , we need to compensate for this by utilizing a composite‐pulse sequence (87) instead of a single blue sideband pulse. Up to an overall phase factor, this transformation yields the desired controlled phase gate. The sequence is composed of four blue sideband pulses and can be described by

24.15 equation

For an intuitive picture of images we plot the evolution of the Bloch vector in Figure 24.9 01. This phase gate is transformed into a controlled‐NOT operation if sandwiched between two images ‐carrier pulses on the target ion, images .

We realize this gate operation ( 27, 88) with a sequence of laser pulses. A blue sideband images ‐pulse, images , on the control ion transfers its quantum state to the bus mode. Then we apply the controlled‐NOT operation images to the target ion. Finally, the bus mode and the control ion are reset to their initial states by another images ‐pulse images on the blue sideband. The gate reaches a fidelity02 of 0.71 ± 0.03 ( 27, 88). If the control qubit is initialized in a superposition state images and the target qubit in images , the controlled‐NOT operation generates an entangled state images . Later, the gate operation was improved and a full process tomography (89) was carried out, yielding a process fidelity of 0.926 ± 0.006.

24.4.5.3 The Sørensen–Mølmer Gate Scheme

Mølmer and Sørensen (9092), and in a different formulation Milburn (93), proposed a gate scheme which does not require perfect ground state cooling. Instead, only the cooling of the ion crystal into the Lamb‐Dicke regime is necessary, such that images 1, where images is the mean thermal phonon number and images the Lamb‐Dicke factor. The authors assume an even number images of ions, which are homogeneously illuminated by a bichromatic laser field with laser frequencies of opposite detunings with respect to the red and blue sideband frequencies, that is, images (see Figure 24.10). The initially prepared state images undergoes a sinusoidal Rabi oscillation to images at an effective Rabi frequency images . In the weak excitation regime images , intermediate levels with vibrational numbers other than images , that is, images and images , are not populated. The effective Rabi frequency reads images . It appears that the ions only absorb photons simultaneously from the bichromatic laser field as the ions share the same vibrational mode. While the absorption of a single photon is suppressed due to the frequency mismatch images , the coupling of the ions to the common vibration mode allows a mutual compensation of this frequency mismatch. As a consequence, the Sørensen–Mølmer scheme works with any even number of ions in a string. For two ions, an effective spin‐spin interaction images is realized, which is an entangling interaction. If the gate evolution is stopped at images , one has generated an entangled state of the electronic components of the ions only, while the state of the vibrational mode is disentangled from the qubit state. This results in a unitary transformation

24.16 equation
Scheme for Level scheme for the Sϕrensen-Mϕlmer gate operation.

Figure 24.10 Level scheme for the Sørensen–Mølmer gate operation. The bichromatic laser field is resonant to the sideband (dashed) of the quantum bus mode and couples qubit states of two ions. Transitions between images and images are driven from this two‐photon resonant process. For a chosen duration this yields the entangled state of images and images . Population in images and images are suppressed because of the frequency detuning and of a quantum interference of two excitation amplitudes.

Image described by caption and surrounding text.

Figure 24.11 Calculation of the evolution of the probabilities for finding both ions in the ground state images and both in the excited state images upon bichromatic driving of the Simages Dimages transition of images Caimages . The parameters are images  = 700 kHz, images 630 kHz (w.r.t. the carrier transition), a mean thermal number of images , and a Lamb‐Dicke parameter of images  = 4.3%. (a) For a Rabi frequency of images 44 kHz the entangled state Eq. 24.16 is generated after an interaction time of 3 ms (0.27 ms). Residual off‐resonant excitation of the intermediate states of odd parity, images and images is shown in gray. (b) For the higher Rabi frequency of 177 kHz the gate operation is performed after 0.225 ms. The amplitude of the off‐resonant excitations is higher, but these vanish exactly at the gate time, when the populations in images and images are balanced at 0.5.

The two ions are in an entangled state after the laser pulse. However, the weak excitation regime implies a low Rabi frequency, and the evolution at images becomes correspondingly slow. This leads impractically long gate durations of typically a few milliseconds. To illustrate this, a calculation of the dynamics of the evolution of the populations images and images according to (92) is displayed in Figure 24.11.

The scheme was significantly improved shortly after the initial proposal, when Sørensen and Mølmer realized that the gate operation could also be driven much faster. However, with faster evolution of images and images and a larger images , intermediate levels with vibrational numbers images are populated, and in general the vibrational quantum number no longer remains unaffected by the evolution. The internal electronic states are therefore entangled with the ion motion during the course of the gate operation. For a successful gate operation, we have to make sure that the vibrational mode returns back to its initial state at the end of the gate operation. This corresponds to a closed circle in the phase space of the gate mode. As shown in (91, 92), the interaction time images has to be adjusted to fulfill images with images . Figure 24.12a) shows the measured population evolution of such a fast bichromatic two‐ion entangling gate. Compared to the simulation in Figure 24.11, the Rabi frequency images is increased, which allows the generation of an entangled state after 0.27 ms. Effects that limit the gate fidelity are discussed in (92). The fast Sørensen–Mølmer entanglement operation was realized for images Caimages ions with a fidelity of 0.993 ± 0.003 (94) on the optical qubit, see Figure 24.12. Further refinement of the method, for example, by shaping the bichromatic laser field, is outlined in (95), and the operation for thermally excited ions even at Doppler cooling temperatures was demonstrated (96). Recently, bichromatic gates have been demonstrated on images Beimages ions at a gate error as low as images by the NIST ion trapping group (43).

Scheme for Realization of the Sϕrensen-Mϕlmer gate operation with a pair of 40Ca+ ions.

Figure 24.12 Realization of the Sørensen–Mølmer gate operation with a pair of images Caimages ions. (a) Evolution of the populations pimages , pimages , and pimages and pimages induced by a bichromatic pulse of duration images . The Rabi frequency images is adiabatically switched on and off within 2 images s and adjusted such that a maximally entangled state is created at images 50 images s. The dashed lines are calculated for mean phonon number of 0.05, neglecting pulse shaping and off‐resonant carrier excitation. The solid lines are obtained by numerically solving the Schrödinger equation for time‐dependent images and imbalanced Rabi frequencies images 1.094. (b) A images analysis pulse applied to both ions prepared in images gives rise to a parity oscillation images as a function of the analysis pulse phase images , where the parity contrast images 0.990(1) is a measure for the gate fidelity (94).

24.4.5.4 The images Geometric Phase Gate

Holonomic quantum computing has been discussed in Section V D, and a proposal for single and two‐qubit phase gates exists for the case of trapped ions (97). Here, depending on the global state configuration of a set of ions, ion motion is transiently excited such that the corresponding trajectories in phase space are closed. This yields a state‐dependent Berry phase given by area enclosed by the trajectories. Thus, these gate operations are robust, since small variations of the actual path from the desired one do affect the accumulated Berry phase to first order.

Image described by caption and surrounding text.

Figure 24.13 Geometric phase gate: (a) beam geometry: laser beams R1 and R2 cross under an angle such that a standing wave pattern along the trap images ‐axis is formed. (b) The resulting action of the state depended optical dipole force onto the breathing mode of a two ion crystal is sketched for all computational basis states 1–4. (c) The resulting trajectories for the breathing mode in phase space, which undergoes transient excitation for states images and images , but not for the even parity states images and images . This leads to the accumulation of a state‐dependent geometric phase.

Experimentally, such a geometric two‐ion geometric phase gate has been realized first by the NIST group (98). Two images Beimages ions are held in a linear trap and are exposed to off‐resonant laser beams (see Figure 24.7) each at an angle of 45images with respect to the trap axis, such that the resulting difference vector images points along the axial direction (98). Up to a small detuning images , the frequency difference of both Raman beams is set close to the breathing mode frequency. For appropriate choice of the laser polarizations, this gives rise to alternating optical polarization along the trap axis. This in turn causes an ac Stark shift which oscillates in space along the trap axis and in time, near the breathing mode frequency. This leads to a near‐resonant optical dipole force, which can coherently excite breathing mode oscillations. The sign and magnitude of the force depends on the global spin configuration, as expressed by the parity operator images (see Figure 24.13). If the interion distance images is chosen to fulfill images with integer images , the optical field at the positions of both ions is identical, and the optical force acts on both ions in opposite directions if the spin configuration is odd, that is, for the states images and images . For even configurations, images and images , the forces on both ions act in the same direction and the breathing mode is not excited. Hence, we obtain excitation of breathing mode oscillation only for odd state configurations. As the difference frequency of both driving beams is deviates from the mode frequency by a detuning images , the force becomes out‐of‐phase with the oscillation and counteracts it after some time. Therefore, the breathing mode oscillation is of transient nature and vanishes after the gate time images . This is analogous to a classical pendulum which is excited slightly off resonance. Its oscillation amplitude initially grows, and the phase difference between external drive and pendulum eigenfrequency is accumulated, such that it is de‐excited again. As a consequence the phase space trajectory is a closed circle for odd state configurations. This leads to accumulation of a state‐dependent geometric phase, which is given by the phase‐space area enclosed by the respective trajectories. If, for a given detuning images , the force magnitude is adjusted such that the enclosed area corresponds to a differential phase of images , this realizes a unitary transform corresponding to a controlled‐phase gate:

24.17 equation

The gate operation does not require perfect ground state cooling, as the trajectories in phase space close regardless of their initial vibrational quantum states, which leads to the required final disentanglement between qubit state and motional state. The gate speed – in contrast to the Cirac–Zoller gate scheme – is not limited by off‐resonant carrier excitations (99). Today, measured gate fidelities reach 0.999(1) (42) and are limited mainly by the spontaneous photon scattering. Increasing the detuning images of the gate drive fields from the corresponding atomic transition leads to loss of drive strength scaling with images , but suppression of photon scattering with images . Therefore, increasing both the drive laser power and the detuning can yields increased gate fidelity.

Qubit operations as this geometric or the Mölmer & Sörensen gate, see Section 24.4.5.6, are ideal in combination with the quantum CCD architecture which is used in segmented micro traps, and which requires a high overhead in ion shuttling operations, see Section 24.4.2. In order to mitigate any effect of residual motional excitation of the phonon number in direction of the trap axis, thus in images ‐direction, we have implemented the geometric gate on the radial rocking mode such that a high fidelity is maintained even for extended algorithms with a few hundred ion reconfiguration operations (100).

24.4.5.5 Pulsed Ultra‐Fast Gates

While frequency combs based on mode‐locked pulsed lasers have become a reliable tool for atomic clocks and optical frequency standards, and their applicability for trapped ion quantum computing is currently explored. For a discussion of the first proposal by Garcia‐Ripoll et al. (101), based on state‐dependent geometric phases.

Short pulses allow for high peak pulse power and correspondingly large Rabi frequencies, and thus for quantum gates of substantially reduced durations. For short pulses, the laser–ion interaction leads to an impulsive and spin‐dependent momentum kick when images exceeds the vibrational trap frequency images , such that quantum logic operations can be performed within a fraction of the vibrational trap period. The convenient choice is a frequency tripled Nd:YVOimages laser at 355 nm, with drives stimulated Raman transitions in images Ybimages ions, tuned halfway between and far off‐resonant to the Simages Pimages and Simages Pimages fine structure transitions. This requires that the bandwidth of the pulses is sufficiently narrow to avoid resonant driving of the dipole transitions. It turns out that the frequency tripled Nd:YVOimages laser offers an optimum trade‐off between undesired scattering‐induced decoherence and the desired large spin‐dependent light shift. Depending on the details of the ion level system, typical pulse durations images between 0.5 and 25 ps are required, such that bandwidth images is much smaller than the fine‐structure splitting images .

Electro‐optical pulse pickers and controlled delay lines serve to tailor gate pulse sequences from the pulse train emanated by the laser source. From the resulting sequence of momentum‐kicks, a closed trajectory in phase space is achieved and results in a spin‐dependent quantum phase ( 38,102). Spin‐motion entanglement has been controlled in this way for a single ion within less that 3 ns (103) and Schrödinger cat states have been generated within 14 ns with a fidelity of 0.88  ±  0.02 (104).

24.4.5.6 The Mintert–Wunderlich Gate

All gate schemes presented above rely on the controlled transfer of momentum from a laser field to the vibration bus mode of the ion crystal, which is governed by the Lamb‐Dicke parameter images . Transitions between hyperfine or Zeeman sublevels of electronic states can be directly driven by rf or microwave fields. As this long‐wavelength radiation displays vanishing values of images and therefore images , we expect no direct momentum transfer. Then, no controlled coupling to vibrational modes would be possible, which precludes two‐qubit gate operations driven by such fields. However, if a magnetic field gradient is applied across the ion crystal, a state‐dependent potential is realized. Then, ions may be transferred between low‐field seeking and high‐field seeking states by a microwave or rf pulse, upon which spatial rearrangement to a different minimum‐energy configuration takes place. As the ions comprising the crystal are rigidly coupled via the Coulomb repulsion, normal modes of vibration can be excited by long‐wavelength radiation. This technique has been coined magnetic gradient induced coupling (MAGIC) (105). A high‐magnetic field gradient can also be harnessed for qubit addressing in frequency space, by tuning the drive frequency to the position‐dependent frequency of one particular qubit. The main advantage of this microwave‐based approach is the high frequency stability and low maintenance effort of commercial off‐the‐shelf microwave sources.

Experimentally, the technique has been demonstrated utilizing either oscillating or with static magnetic field gradients. In the latter case, single qubit addressing was accomplished with a residual crosstalk as small as images (106) and three‐ion entanglement with a fidelity of 0.57 ± 0.04 (34). More recently, the fidelity of magnetic‐gradient enabled two‐qubit entanglement was increased to 0.985 ± 0.012 by utilizing of a segmented microtrap with a magnetic field gradient of 23.6 T/m, where dynamical decoupling was used to mitigate noise (37). For oscillating near‐field microwave driving, an entangling gate fidelity of 0.76 ± 0.03 was reached (35). To generate strong field gradients on the order of 35 T/m, it is required to use surface electrode traps with integrated microwave electrodes. Sufficient coupling is realized for relatively small distances of the ions to the surface of about 30 images m, such that surface‐induced anomalous heating of the ion motion becomes the dominant error source (107). Using long‐lived hyperfine qubits encoded in images Caimages and a microwave‐driven version of the Sørensen–Mølmer scheme, a fidelity of 0.997 ± 0.001 was achieved recently (36). Again, the dominant error source is heating of the ion at 75 images m distance above the gold surface.

24.4.5.7 Quantum Computing Architectures

A future large‐scale universal quantum computer will necessarily rely on quantum error correction, which increases the number of required ions for redundant qubit storage and operations. Additionally, ancilla ions for readout of error syndromes are needed, such that in total on the order of images images ions may be necessary to demonstrate an actual quantum supremacy. This leads to the question of how an architecture hosting a sufficient number of ions can be technologically realized, while retaining excellent control.

The quantum CCD approach (46) is the first attempt to address this question. While in this proposal, gate operations are driven by laser pulses, an interesting alternative would be using microwave pulses (108). However, in both cases we face a large overhead of ion shuttling operations, such that it may be of interest to investigate different means to couple qubits which are stored at different sites. Here, ion–photon interfaces which have the potential to bridge large distances within one quantum processing unit or even between spatially separated processors (109,110). This approach was stimulated by cavity‐QED experiments (111114). Even today, realizing strong coupling of a single ion to an optical cavity mode remains technically challenging. Recently, however, such interfaces have been largely improved and lead to controlled photon–ion entanglement and ion–ion entanglement (115117). Another option for realizing large scale QC is the electric coupling either via antenna structures (118) or directly. Such coupling has been demonstrated for small ion crystals (119,120) and stimulated some further experimental investigations for two‐dimensional arrays of ions (121,122). Obviously, and for all cases, the challenge lies in the manufacturing and control of such complex architectures.

Image described by caption and surrounding text.

Figure 24.14 (a) Space–time diagram of the quantum teleportation algorithm. A Bell state is distributed among Alice (lower part of figure) and Bob. Alice transmits the outcome of her Bell analysis to Bob who recovers the original quantum state. (b) Protocol for teleportation from ion images 1 to ion images 3: Initially, a Bell state of ions images 2 and images 3 is prepared. The state images is encoded in ion images 1. The Bell state analyzer consists of a controlled Z‐gate followed by images /2 rotations and a state detection of ions images 1 and images 2. Note that this implementation uses a Bell basis rotated by images /4 with respect to the standard notation. Therefore, a images /2 rotation on ion images 3 is required prior to the reconstruction operations Z and X. Gray lines indicate qubits which are protected against light scattering. Ions images 1 and images 2 are detected by observing their fluorescence on a PMT. Only upon a detection event images the corresponding reconstruction operation is applied to ion images 3. Classical information is represented by double lines. For the fidelity analysis, we apply images and measure the quantum state of ion images 3.

24.4.6 Quantum Teleportation

In 2004, more than a decade after the proposal (123), deterministic quantum teleportation with matter has been demonstrated in two different laboratories ( 29,30). For the teleportation with light, we refer the reader to Section IV. The quantum teleportation algorithm for a qubit images from Alice to Bob is based on five consecutive steps:

  1. An entangled pair of qubits images is created and distributed, the particle with index 2 to Alice and the other one to Bob.
  2. Alice holds a single qubit in an unknown quantum state images , which is to be teleported to Bob. The overall quantum state of the three qubit system reads images .

    The two particles with index 1 and 2 belong to Alice, and she performs the images ‐gate and the images operation (see Figure 24.13(b)) to rotate the basis for her qubits from the computational basis states images into the Bell basis states denoted by images and images .

  3. The overall quantum state can be written as a sum of four terms
    24.18 equation

    Alice performs a measurement in the Bell basis on both her qubits, particles 1 and 2. Thus, she projects her particles of images into one of the possible Bell states, each with 1/4 probability outcome. If Alice's measurement result is images , Bob's qubit is already in the desired quantum state images .

    If she finds a different outcome, for example, images , Bob has recovered the quantum state images which can be transformed into images by a single qubit rotation images in his particle with index 3. Using the Pauli operators images which generate rotations about the different axes of the Bloch sphere we obtain,

    24.19 equation

    with images and images . Note, that only images and images rotations are necessary.

  4. Therefore, Alice sends the outcome of her measurement, two bits of classical information to Bob, such that
  5. Bob is able to perform the correct single qubit rotation on his particle to obtain back the original state images . Note, that no quantum information is duplicated but that the entire information about the state images has fully disappeared at Alice's side.

The protocol, see Figure 24.14 (b), has been realized (30) with three ions in the linear trap shown in Figure 24.2a). The quantum algorithms consist of more than 30 laser pulses. The outcome of the deterministic teleportation is revealed by inverse reconstruction and shows a fidelity of 75images (30). Subsequently, this result has been improved to 83%, and the analysis of the algorithm has been completed by applying process tomography on the teleported output state ( 25 124126). Teleportation with a fidelity of 78images was demonstrated with three ions in a linear segmented ion trap by the NIST group (29). For quantum gate operations among two ions, for single qubit rotations, and for the readout of single qubits without affecting the other qubits, the required ions are singled out from the linear crystal via ion shuttling, and transported into a processor section of the ion trap where the laser‐ion interactions are driven. Thus, quantum teleportation has become the first algorithm to demonstrate the benefits of segmented and miniaturized Paul traps.

24.4.7 Selected Recent Highlights

As the invention of Shor's factoring algorithm fostered the development of quantum computing, it is intriguing to the progress on its actual realization, recently performed for the case images and based on scalable algorithmic building blocks (39). The problem of finding the prime factors of an integer images can be mapped to the problem of finding the period images for integers images , which is the smallest integer for which modular exponentiation yields zero, that is, images From the resulting period images , one determines prime factors of images as the greatest common divisor of images and images . In the experiment (39), images has been factorized by using a linear ion crystal with five ions in the trap Figure 24.2. It is crucial that the algorithm has been demonstrated without using precompiled gate sequences for the modular exponentiation operation. One ancilla and four register qubits are required for the modular exponentiation images for a suitable base images , and images , to find its period. The measurement results for base images clearly show a period images with a squared statistical overlap larger than 0.90 for all cases. This is the outcome of a large number of experimental repetitions. However, Shor's algorithm should allow for deducing the period with high probability already from a single‐shot measurement. In the experiment this is achieved for all images with a probability greater than 0.5, therefore, after only eight single‐shot repetitions, one would reveal the correct periodicity at a probability exceeding 0.99 (39).

Remarkable progress toward fault‐tolerant quantum computing has been realized by the demonstration of a topologically encoded qubit (40). One logical qubit is encoded in seven physical trapped‐ion qubits. The corresponding stabilizer operators images and images are defined on three different subsets images (plaquettes), each consisting of four ions. State preparation via plaquette‐wise entangling gates and stabilizer readout operations are realized by a global laser beam, such that it is required to hide the other ions from the laser interaction. A tightly focused laser beam is used to transfer these into the metastable Dimages state, such that out of the static crystal of seven ions, only the required sub‐set of qubits is affected by the gate operation.

The technique of individual single ion addressing with a tightly focused laser beam acting on a static linear ion crystal is also used to implement a universal programmable quantum processor (127). All trapped ion qubits are fully connected by the common mode of vibration such that pairwise entangling gates between all possible pairs are realized, which yields a crucial advantage over QC architectures with nearest‐neighbor or star‐shaped coupling topology. Gate operations reach a fidelity of images 0.98. A quantum Fourier transformation performed with 15 gate operations on the five qubits and achieves an average fidelity of 0.619 ± 0.005s. Note, that such static ion crystal approach has been used for quantum simulation of interacting spin systems (128)

Such highlights obtained with static linear ion qubit registers are complemented by experiments toward a scalable re‐configurable quantum CCD. The first demonstration of a complete methods set comprises individual addressed single qubit preparation and state readout, single and two‐qubit laser‐driven gate operations, in combination with a spatial separation of two‐ion crystals, and their recombination as well as transport operations for single ions and small crystals (41), 129. Process tomography reveals a fidelity of 0.987 ± 0.003. The set of operations was completed by the ion‐SWAP operation for two‐ and three‐ion crystals, where a mean process fidelity of 0.995 ± 0.005 is obtained. Only in this way, any of the ions in the linear arrangement can be coupled with any other ion, regardless of the initial positions (130), without using complicated images ‐ or images ‐junctions in segmented traps, see Figure 24.2. Even sequences with about 250 shuttling on a four ion qubit register, and several laser‐driven geometric phase gate operations have been realized (100) and multipartite entanglement with fidelity of about images is achieved.

For algorithms with even higher complexity and thus even more overhead in terms of register reconfigurations operations, the technique of sympathetic cooling is required: An auxiliary ion of a different species is used for ground state cooling of the relevant collective vibrational modes of oscillation of the mixed crystal, thus cooling also the qubit ion(s) without affecting stored quantum states. This way, the quantum gate operations on the reconfigured qubit at persistent high fidelity are enabled. Mixed ion crystals can serve for another crucial purpose: Quantum error correction relies on the readout of error syndromes, but obviously this has to be performed without affecting the stored quantum information. As any readout scheme relies on state‐dependent resonance fluorescence, it is hardly avoidable in a realistic setting that resonant scattered light affects memory qubits in an undesired way – unless a different species with different resonance wavelengths is employed for readout. As an important step toward quantum non‐demolition stabilizer readout, multi‐element logic gates for trapped ion qubits have been realized for mixed crystals of images Beimages and images Mgimages (131), of images Caimages and images Caimages (132) or of images Caimages and images Srimages . Furthermore, quantum logic based optical frequency standards, see Section 36, rely on similar interspecies entanglement gate operations.

Observing the rapid and substantial progress of ion trap quantum computing, overcoming technological barriers, and controlling the required complex experimental environments increasingly well, we are confident that the concepts for scalability to a large number of qubits are sustainable. Trapped ions stay among the most promising platforms for experimental quantum computing.

Acknowledgements

We acknowledge support by Alexander Stahl and Thomas Ruster.

References

  1. 1 Paul, W., Osberghaus, Q., and Fischer, E. (1958) Forschungsberichte des Wirtschafts– und Verkehrsministeriums Nordrhein‐Westfalen, vol. 415.
  2. 2 http://www.nobelprize.org/nobel_prizes/physics/laureates/1989/paul‐lecture.html.
  3. 3 Hänsch, T.W. and Schawlow, A.L. (1975) Opt. Commun., 13, 68.
  4. 4 Wineland, D. and Dehmelt, H. (1975) Bull. Am. Phys. Soc., 20, 637–637.
  5. 5 Neuhauser, W., Hohenstatt, M., Toschek, P.E., and Dehmelt, H. (1980) Phys. Rev. A, 22, 1137.
  6. 6 Nagourney, W., Sandberg, J., and Dehmelt, H. (1986) Phys. Rev. Lett., 56, 2797.
  7. 7 Sauter, T., Neuhauser, W., Blatt, R., and Toschek, P.E. (1986) Phys. Rev. Lett, 57, 1696.
  8. 8 Schubert, M., Siemers, I., Blatt, R., Neuhauser, W., and Toschek, P.E. (1995) Phys. Rev. A, 52, 2994.
  9. 9 Rempe, G., Schmidt‐Kaler, F., and Walther, H. (1990) Phys. Rev. Lett., 64, 2783.
  10. 10 Brune, M. et al. (1996) Phys. Rev. Lett., 76, 1800.
  11. 11 Raimond, J.‐M., and Haroche, S. (2006) Exploring the Quantum: Atoms, Cavities, and Photons, Oxford University Press.
  12. 12 Meekhof, D.M., Monroe, C., King, B.E., Itano, W.M., and Wineland, D.J. (1996) Phys. Rev. Lett., 76, 1796.
  13. 13 Diedrich, F., Bergquist, J., Itano, W., and Wineland, D.J. (1989) Phys. Rev. Lett., 62, 403.
  14. 14 Monroe, C., Meekhof, D., King, B., and Wineland, D. (1996) Science, 272, 1131.
  15. 15 Gleyzes, S. et al. (2007) Nature, 446, 297.
  16. 16 https://www.nobelprize.org/nobel_prizes/ physics/laureates/2012/advanced.html.
  17. 17 Manin, Yu.I. (1980) Computable and uncomputable (in Russian), Sovetskoye Radio, Moscow.
  18. 18 Manin, Yu.I. (1999) Nordrhein‐Westfael. Akademie der Wissenschaften, arXiv:quant‐ph/9903008.
  19. 19 Feynman, R. (1982) Int. J. Theor. Phys., 21, 467.
  20. 20 Feynman, R. (1986) Found. Phys., 21, 507.
  21. 21 Deutsch, D. (1989) Proc. R. Soc. London, Ser. A, 425, 73.
  22. 22 Shor, P.W. and Goldwasser, S. (1994) Algorithms for quantum computation: discrete logarithms and factoring. Proceedings of the 35th Annual Symposium on the Foundations of Computer Science.
  23. 23 Ekert, A. (1994) At. Phys., 14, 450.
  24. 24 Ekert, A. and Josza, R. (1996) Rev. Mod. Phys., 68 (3), 733.
  25. 25 Nielsen, M.A. and Chuang, I.L. (2000) Quantum Computation and Quantum Information, Cambridge University Press.
  26. 26 Monroe, C. et al. (1995) Phys. Rev. Lett., 75, 4011.
  27. 27 Schmidt‐Kaler, F. et al. (2003) Appl. Phys. B, 77, 789.
  28. 28 Haroche, S. and Raimond, J.M. (1996) Phys. Today, 49 (8), 51.
  29. 29 Barrett, M.D. et al. (2004) Nature, 429, 737.
  30. 30 Riebe, M. et al. (2004) Nature, 429, 734.
  31. 31 Riebe, M. et al. (2007) New J. Phys., 9, 211.
  32. 32 Rowe, M. et al. (2001) Nature, 409, 791.
  33. 33 Monz, T. et al. (2011) Phys. Rev. Lett., 106, 130506.
  34. 34 Khromova, A. et al. (2012) Phys. Rev. Lett., 108, 220502.
  35. 35 Ospelkaus, C. et al. (2011) Nature, 476 (7359), 181.
  36. 36 Harty, T.P. et al. (2016) Phys. Rev. Lett., 117, 140501.
  37. 37 Weidt, S. et al. (2016) Phys. Rev. Lett., 117, 220501.
  38. 38 Campbell, W.C. et al. (2010) Phys. Rev. Lett., 105, 090502.
  39. 39 Monz, T., Nigg, D., Martinez, E.A., Brandl, M.F., Schindler, P., Rines, R., Wang, S.X., Chuang, I.L., and Blatt, R. (2016) Science, 351, 1068.
  40. 40 Nigg, D. et al. (2014) Science, 345, 302.
  41. 41 Home, J.P. et al. (2009) Science, 325, 1227.
  42. 42 Ballance, C.J., Harty, T.P., Linke, N.M., Sepiol, M.A., and Lucas, D.M. (2016) Phys. Rev. Lett., 117, 060504.
  43. 43 Gaebler, J.P. et al. (2016) Phys. Rev. Lett., 117, 060505.
  44. 44 Dehmelt, H. (1975) Bull. Am. Phys. Soc., 20, 60.
  45. 45 Harty, T.P. et al. (2014) Phys. Rev. Lett., 113, 220501.
  46. 46 Kielpinski, D., Monroe, C., and Wineland, D.J. (2002) Nature, 417, 709.
  47. 47 Ghosh, P.K. (1995) Ion Traps, The International Series of Monographs on Physics, vol. 90, Oxford Science Publications.
  48. 48 Gulde, S. et al. (2003) Nature, 421, 48.
  49. 49 Drewsen, M. and Broner, A. (2000) Phys. Rev. A, 62, 045401.
  50. 50 Nägerl, H.C., Becher, W., Eschner, J., Schmidt‐Kaler, F., and Blatt, R. (1998) Appl. Phys. B, 66, 603.
  51. 51 Waki, I., Kassner, S., Birkl, G., and Walther, H. (1992) Phys. Rev. Lett., 68, 2007.
  52. 52 Birkl, G., Kassner, S., and Walther, H. (1992) Nature, 357, 310.
  53. 53 Raizen, M.G. et al. (1992) Phys. Rev. A, 45, 6493.
  54. 54 Nägerl, H.C., Bechter, W., Eschner, J., Schmidt‐Kaler, F., and Blatt, R. (1998) Opt. Express, 3, 89.
  55. 55 James, D.V.F. (1998) Appl. Phys., B66, 181.
  56. 56 Steane, A. (1997) Appl. Phys. B, 64, 632.
  57. 57 Marquet, C., Schmidt‐Kaler, F., and James, D.F.V. (2003) Appl. Phys. B, 76, 199.
  58. 58 Enzer, D.G. et al. (2000) Phys. Rev. Lett., 85, 2466.
  59. 59 Brownnutt, M., Kumph, M., Rabl, P., and Blatt, R. (2015) Rev. Mod. Phys., 87, 1419.
  60. 60 Poitzsch, M.E., Bergquist, J.C., Itano, W.M., and Wineland, D.J. (1995) Rev. Sci. Instrum., 67, 129.
  61. 61 Labaziewicz, J., Ge, Y., Leibrandt, D.R., Wang, S.X., Shewmon, R., and Chuang, I.L. (2008) Phys. Rev. Lett., 101, 180602.
  62. 62 Brandl, M.F. et al. (2016) Rev. Sci. Instrum., 87 (11), 113103.
  63. 63 Hite, D.A. et al. (2012) Phys. Rev. Lett., 109, 103001.
  64. 64 Daniilidis, N. et al. (2014) Phys. Rev. B, 89, 245435.
  65. 65 Stick, D. et al. (2006) Nat. Phys., 2, 36.
  66. 66 Britton, J., Leibfried, D., Beall, J.A., Blakestad, R.B., Wesenberg, J.H., and Wineland, D.J. (2009) Appl. Phys. Lett., 95 (17), 173102.
  67. 67 Wilpers, G., See, P., Gill, P., and Sinclair, A.G. (2012) Nat. Nano, 7 (9), 572.
  68. 68 Seidelin, S. et al. (2006) Phys. Rev. Lett., 96, 253003.
  69. 69 Amini, J.M. et al. (2010) New J. Phys., 12 (3), 033031.
  70. 70 Bowler, R., Warring, U., Britton, J.W., Sawyer, B.C., and Amini, J. (2013) Rev. Sci. Instrum., 84 (3), 033108.
  71. 71 Baig, M.T., Johanning, M., Wiese, A., Heidbrink, S., Ziolkowski, M., and Wunderlich, C. (2013) Rev. Sci. Instrum., 84 (12), 124701.
  72. 72 Rowe, M.A. et al. (2002) Quantum Inf. Comput., 2, 257.
  73. 73 Bowler, R. et al. (2012) Phys. Rev. Lett., 109, 080502.
  74. 74 Walther, A. et al. (2012) Phys. Rev. Lett., 109, 080501.
  75. 75 Ruster, T. et al. (2014) Phys. Rev. A, 90, 033410.
  76. 76 Kaufmann, H., Ruster, T., Schmiegelow, C.T., Schmidt‐Kaler, F., and Poschinger, U.G. (2014) New J. Phys., 16 (7), 073012.
  77. 77 Leibfried, D., Blatt, R., Monroe, C., and Wineland, D. (2003) Rev. Mod. Phys., 75, 281.
  78. 78 Haeffner, H., Roos, C., and Blatt, R. (2008) Phys. Rep., 469, 155.
  79. 79 Benhelm, J., Kirchmair, G., Roos, C.F., and Blatt, R. (2008) Nat. Phys., 4 (6), 463.
  80. 80 Ozeri, R. et al. (2005) Phys. Rev. Lett., 95, 030403.
  81. 81 Ruster, T. et al. (2016) Appl. Phys. B, 122 (10), 254.
  82. 82 Cirac, I. and Zoller, P. (1995) Phys. Rev. Lett., 74, 4091.
  83. 83 Roos, Ch. et al. (1999) Phys. Rev. Lett.. 83, 4713.
  84. 84 Roos, C.F. et al. (2000) Phys. Rev. Lett., 85, 5547.
  85. 85 Barrett, M. et al. (2003) Phys. Rev. A, 68, 042302.
  86. 86 Nägerl, H.C. et al. (1999) Phys. Rev. A, 60, 145.
  87. 87 Cilds, A.M. and Chuang, I.L. (2000) Phys. Rev. A, 63, 012306.
  88. 88 Schmidt‐Kaler, F. et al. (2003) Nature, 422, 408.
  89. 89 Riebe, M. et al. (2006) Phys. Rev. Lett., 97, 220407.
  90. 90 Mølmer, K. and Sørensen, A. (1999) Phys. Rev. Lett., 82, 1835.
  91. 91 Sørensen, A. and Mølmer, K. (1999) Phys. Rev. Lett., 82, 1971.
  92. 92 Sørensen, A. and Mølmer, K. (2000) Entanglement and quantum computation with ions in thermal motion, arXiv:quant‐ph/0002024.
  93. 93 Milburn, G.J. (1999) Simulating nonlinear spin models in an ion trap, arXiv:quant‐ph/9908037.
  94. 94 Benhelm, J., Kirchmair, G., Roos, C.F., and Blatt, R. (2008) Nat. Phys., 4, 463.
  95. 95 Roos, C.F. (2008) New J. Phys., 10, 013002.
  96. 96 Kirchmair, G. et al. (2009) New J. Phys., 11, 023002.
  97. 97 Duan, L.M., Cirac, J.I., and Zoller, P. (2001) Science, 292, 1695.
  98. 98 Leibfried, D. et al. (2003) Nature, 422, 412.
  99. 99 Steane, A. et al. (2000) Phys. Rev. A, 62, 042305.
  100. 100 Kaufmann, H. et al. (2017) Phys. Rev. Lett., 119, 150503.
  101. 101 Garcia‐Ripoll, J.J., Zoller, P., and Cirac, J.I. (2003) Phys. Rev. Lett., 91, 157901.
  102. 102 Mizrahi, J. et al. (2014) Appl. Phys. B, 114, 45.
  103. 103 Mizrahi, J. et al. (2013) Phys. Rev. Lett., 110, 203001.
  104. 104 Johnson, K.G., Wong‐Campos, J.D., Neyenhuis, B., Mizrahi, J., and Monroe, C. (2017) Nat. Comm., 8, 697.
  105. 105 Mintert, F. and Wunderlich, C. (2001) Phys. Rev. Lett., 87, 257904.
  106. 106 Piltz, Ch., Sriarunothai, Th., Varón, A.F., and Wunderlich, Ch. (2014) Nat. Commun., 5, 4679.
  107. 107 Warring, U. et al. (2013) Phys. Rev. A, 87, 013437.
  108. 108 Lekitsch, B. et al. (2017) Sci. Adv., 3, e1601540.
  109. 109 Monroe, C. et al. (2014) Phys. Rev. A, 89, 022317.
  110. 110 Monroe, C. and Kim, J. (2013) Science, 339, 1164.
  111. 111 Kreuter, A. et al. (2004) Phys. Rev. Lett., 92, 203002.
  112. 112 Mundt, A.B. et al. (2002) Phys. Rev. Lett., 89, 103001.
  113. 113 Guthöhrlein, G.R., Keller, M., Hayasaka, K., Lange, W., and Walther, H. (2001) Nature, 414, 49.
  114. 114 Keller, M., Lange, B., Hayasaka, K., Lange, W., and Walther, H. (2004) Nature, 431, 1075.
  115. 115 Duan, L.‐M. and Monroe, C. (2010) Rev. Mod. Phys., 82, 1209.
  116. 116 Kurz, C. et al. (2014) Nat. Commun., 5, 5527.
  117. 117 Casabone, B., Northup, T.E. et al. (2013) Phys. Rev. Lett., 111, 100505.
  118. 118 Daniilidis, N., Lee, T., Clark, R., Narayanan, S., and Häffner, H. (2009) J. Phys. B, 42, 144012.
  119. 119 Harlander, M., Lechner, R., Brownnutt, M., Blatt, R., and Hänsel, W. (2011) Nature, 471, 200.
  120. 120 Jost, J.D. et al. (2011) Nature, 459, 683.
  121. 121 Kumph, M., Brownnutt, M., and Blatt, R. (2011) New J. Phys., 13, 073043.
  122. 122 Mielenz, M. et al. (2016) Nat. Commun., 7, 11839.
  123. 123 Bennett, C.H. et al. (1993) Phys. Rev. Lett., 70, 1895.
  124. 124 Roos, C.F. et al. (2004) Phys. Rev. Lett., 92, 220402.
  125. 125 O'Brien, J.L. et al. (2004) Phys. Rev. Lett., 93, 080502.
  126. 126 Hradil, Z. et al. (2004) 3, maximum‐likelihood methods in quantum mechanics, in Quantum State Estimation, Lecture Notes in Physics (eds M.G.A. Paris and J. Rehacek), Springer‐Verlag, p. 59.
  127. 127 Debnath, S. et al. (2016) Nature, 536 (7614), 63.
  128. 128 Blatt, R. and Roos, C.F. (2012) Nat. Phys., 8, 277.
  129. 129 Kaufmann, H. et al. (2017) Phys. Rev. Lett., 119, 150503.
  130. 130 Kaufmann, H. et al. (2017) Phys. Rev. A, 95, 052319.
  131. 131 Tan, T. et al. (2015) Nature, 528, 380.
  132. 132 Ballance, C.J. et al. (2015) Nature, 528, 384.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset