Archive for the ‘Quantum Computing’ Category

Quantum computing gets hardware boost with D-Waves spin glass breakthrough – TechRepublic

One of the challenges in quantum computing is overcoming 3D spin-glass optimization limitations, which can slow down quantum simulation meant to solve real-world optimization problems. An experimental solution is D-Waves Advantage quantum computer, running spin-glass dynamics (essentially a sequence of magnets) on 5,000 qubits.

According to a study by scientists from D-Wave and Boston University, published in the journal Nature, the team has validated that quantum annealing a mathematical process used to find low-energy states by using quantum fluctuations can improve solution quality faster than classical algorithms, at least theoretically. It may be a key step forward in showing the ways in which a quantum processor can compute coherent quantum dynamics in large-scale optimization problems.

D-Wave customers who subscribe to the Leap quantum cloud service can access the new commercial-grade, annealing-based quantum computer as of April 19.

Jump to:

The main takeaway for enterprises is that spin-glass computing on a quantum annealing device may eventually be able to efficiently solve optimization problems, achieving a goal with as little energy as possible. For example, it could be a relatively efficient way to answer questions such as Should I ship this package on this truck or the next one? or the traveling salesman problem (What is the most efficient route a traveling salesperson should take to visit different cities?), as D-Wave wrote.

D-Wave is one of the only companies that offers enterprise quantum computing space with both gate and annealing programs, which now includes its 5,000 qubit, commercial-grade Advantage quantum computer. There is still some question as to how practical this technology is, but the new paper is proof that further commercial quantum computing optimization can be performed on D-Waves hardware.

SEE: Should IT teams factor quantum computing into their decisions?

Getting deeper into the physics, spin glasses are often used as test beds for paradigmatic computing, the researchers said, but using this approach in a programmable system and therefore one that can be used to do practical calculations still leads to potential problems. D-Wave has solved this on its hardware by using quantum-critical spin-glass dynamics on thousands of qubits with a superconducting quantum annealer.

The same hardware that has already provided useful experimental proving ground for quantum critical dynamics can be also employed to seek low-energy states that assist in finding solutions to optimization problems, said Wojciech Zurek, theoretical physicist at Los Alamos National Laboratory and leading authority on quantum theory, in D-Waves press release.

Applications that solve optimization problems like the packaging shipping question above require a minimum energy state from the quantum annealing processors they run on. Other calculations that could be used for decision-making, such as probabilistic sampling problems, need good low-energy samples in order to run.

D-Wave says spin glasses can be brought into low-energy states faster by annealing quantum fluctuations than by conventional thermal annealing.

This paper gives evidence that the quantum dynamics of a dedicated hardware platform are faster than for known classical algorithms to find the preferred, lowest energy state of a spin glass, and so promises to continue to fuel the further development of quantum annealers for dealing with practical problems, said Gabriel Aeppli, professor of physics at ETH Zrich and EPF Lausanne, and head of the Photon Science Division of the Paul Scherrer Institut.

Another problem researchers in the quantum computing world are trying to solve is qubit coherence. In a simplified sense, coherence means that a quantum state maintains certain physical qualities while in use. Research shows that coherent quantum annealing can improve solution quality faster than classical algorithms.

Hand-in-hand development of the gate and annealing programs will bring us to longer coherence times and better qubit parameters, allowing our advantage over classical optimization to grow, Andrew King, director of performance research for D-Wave, wrote in a blog post.

Quantum annealing can be used for a wide range of practical optimization applications, Murray Thom, vice president of quantum business innovation at D-Wave, told TechRepublic. For example, it is already being used today to optimize supply chains, employee scheduling, e-commerce delivery, missile defense, protein folding, fraud detection, and industrial manufacturing, just to name a few.

While the newly published research was conducted on the currently commercially available Advantage quantum computer, D-Wave is also working on its next iteration. The Advantage2 system is in the experimental prototype stage and will be D-Waves sixth-generation quantum computing hardware. D-Wave anticipates the full Advantage2 system will launch with 7,000 qubits and does not have a projected release date for the alpha version.

Moving forward, we expect to see annealing quantum computers increase their business impact as technological enhancements in qubit coherence yield higher quality solutions and enable higher connectivity architectures, Thom said.

He expects over the next five years to see more development of gate model systems, an alternative to quantum annealing which has seen some interest from the pharmaceutical industry.

Scalable, error-corrected gate-model machines will begin to emerge and begin to show early business value, he said.

D-Wave competes with other quantum computing systems and software providers such as Rigetti Computing, Google, IBM, Honeywell, and IonQ.

Go here to read the rest:
Quantum computing gets hardware boost with D-Waves spin glass breakthrough - TechRepublic

Joe Rogan and Michio Kaku Compare Reality to The Matrix as the Duo Address the Wonders of Quantum Computing – EssentiallySports

UFC commentator, Joe Roganhas achieved great success as a podcaster. The biggest reason behind the success of The Joe Rogan Experience is how it entertains people of different interests. Even though Rogan has been an MMA analyst for over two decades, he never made his podcast all about MMA. The 55-year-old invites guests from all walks of life to his show, and that can vary from comedians to scientists. Other than MMA, Rogan has always shown a keen interest in science and technology. He loves to dig deeper into the unknown secrets of the world. Moreover, he is fascinated by the growth of technology over time.

In JRE, the former Fear Factor host has often expressed his concern over the uncontrollable growth of technology. Rogan fears a world where science and technology overpower human efficiency. In a recent episode of JRE, the UFC commentator talked to Dr Michio Kaku. Kaku is a professor of theoretical physics. The two discussed a lot about the tremendous growth of technology over the years.

ADVERTISEMENT

Article continues below this ad

In the #1980 episode of JRE, Rogan and Dr Michio Kaku connected the realities of the world to the movie The Matrix. Kaku had written a book called Quantum Supremacy: How the Quantum Computer Revolution Will Change Everything.His vast knowledge of quantum computers made the UFC commentator eager to learn more.

During the conversation, Rogan asked, But if we are talking about technology as we currently understand it today, in comparison to technology as they had available to them a thousand years ago. What we do now, thats insane? And you are talking about quantum computing, which is almost available today and you look at thousand years from now. Couldnt you potentially imagine there could be a world where there is technology sufficient to do what we are talking about? To create a version of reality?

Kaku was able to give a clear answer to Rogans query. He said, Well, if you saw the movie we are all in parts and we are all connected to computers, then stimulus, The Matrix. As long as you are stimulating a piece of the matrix, not the whole thing, but as you walk from place to place, the computer reassembles and replicates that place. That may be possible, but not the whole earth.

Rogans questions to Kaku proved the amount of interest he has in science and technology.

ADVERTISEMENT

Article continues below this ad

In the very same episode of JRE, Rogan asked Kaku about the quantum computers that have the ability to break any code. Rogan was concerned about the world without secrets. However, Kaku didnt seem worried like Rogan. The professor told him that the secret codes of some nations had been broken in the past as well.

READ MORE: It Was Like an Assassination: Joe Rogan Recalls Mike Tyson at His Destructive Best in His Most Violent Fight

ADVERTISEMENT

Article continues below this ad

He went on to explain about World War II, where the secret codes of the Germans were broken using quantum computers. This helped to know the places where Germany was planning to attack beforehand. Back then, it was technology that helped save millions of lives.

Watch This Story:How Did Joe Rogan Become Famous?

Read the rest here:
Joe Rogan and Michio Kaku Compare Reality to The Matrix as the Duo Address the Wonders of Quantum Computing - EssentiallySports

Non-Abelian braiding of graph vertices in a superconducting processor – Nature.com

Elementary particles in three dimensions are either bosons or fermions. The existence of only two types is rooted in the fact that the worldlines of two particles in three plus one dimensions can always be untied in a trivial manner. Hence, exchanging a pair of indistinguishable particles twice is topologically equivalent to not exchanging them at all, and the wavefunction must remain the same. Representing the exchange as a matrix R acting on the space of wavefunctions with a constant number of particles, it is thus required that R2=1 (a scalar), leaving two possibilities: R=1 (bosons) and R=1 (fermions). Such continuous deformation is not possible in two dimensions, thus allowing collective excitations (quasiparticles) to show richer braiding behaviour. In particular, this permits the existence of Abelian anyons2,3,6,7,8,24,25, in which the global phase change due to braiding can take any value. It has been proposed that there exists another class of quasiparticles known as non-Abelian anyons, in which braiding instead results in a change of the observables of the wavefunction4,5,24. In other words, R2 does not simplify to a scalar, but remains a unitary matrix. Therefore, R2 is a fundamental characteristic of anyon braiding. The topological approach to quantum computation26 aims to leverage these non-Abelian anyons and their topological nature to enable gate operations that are protected against local perturbations and decoherence errors5,27,28,29,30. In solid-state systems, primary candidates of non-Abelian quasiparticles are low-energy excitations in Hamiltonian systems, including the 5/2 fractional quantum Hall states31,32, vortices in topological superconductors33,34 and Majorana zero modes in semiconductors proximitized by superconductors35,36,37,38. However, direct verification of non-Abelian exchange statistics has remained elusive39,40,41.

We formulate the necessary requirements for experimentally certifying a physical system as a platform for topological quantum computation5,26: (1) create an anyon pair; (2) verify the rules that govern the collision of two anyons, known as the fusion rules; (3) verify the non-Abelian braiding statistics reflected in the matrix structure R2 and (4) realize controlled entanglement of anyonic degrees of freedom. Notably, the observation of steps (2)(4) requires measurements of multi-anyon states, by means of fusion or non-local measurements.

The advent of quantum processors allows for controlled unitary evolution and direct access to the wavefunction rather than the parameters of the Hamiltonian. These features enable the use of local operations for efficient preparation of topological states that can host non-Abelian anyons, andas we will demonstratetheir subsequent braiding and fusion. Moreover, these platforms allow for probing arbitrary Pauli strings through destructive multiqubit (that is, non-local) measurements. As the braiding of non-Abelian anyons in this platform is achieved through unitary gate control rather than adiabatic evolution of a Hamiltonian system, we note that the anyons are not quasiparticles in the sense of eigenstates that persist throughout a Hamiltonian evolution. Their movement is achieved through local operations along their paths, and they are kept spatially separated throughout the braiding. We therefore emphasize that the two-dimensional braiding processes are physically taking place on the device, leading to actual non-Abelian exchange effects of local anyons in the many-body wavefunction, rather than matrix operations that simply follow the same algebra.

To realize a many-body quantum state that can host anyons, it is essential to control the topological degeneracy. A suitable platform for achieving this requirement is a stabilizer code42, in which the wavefunctions are characterized by a set of commuting operators ({{hat{S}}_{p}}) called stabilizers, with ({hat{S}}_{p}left|psi rightrangle ={s}_{p}left|psi rightrangle ) and sp=1. The code space is the set of degenerate wavefunctions for which sp=1 for all p. Hence, every independent stabilizer divides the degeneracy of the code space by two.

Whereas the physical layout of qubits is typically used to determine the structure of the stabilizers, the qubits can be considered to be degree j vertices (DjV; j{2,3,4}) on more general planar graphs (Fig. 1a)23. Using this picture, each stabilizer can be associated with a plaquette p, whose vertices are the qubits on which ({hat{S}}_{p}) acts:

$${hat{S}}_{p}=prod _{v,in ,{rm{vertices}}}{hat{tau }}_{p,v}.$$

(1)

({hat{tau }}_{p,v}) is here a single-qubit Pauli operator acting on vertex v, chosen to satisfy a constraint around that vertex (Fig. 1b). An instance where sp=1 on a plaquette is called a plaquette violation. These can be thought of as quasiparticles, which are created and moved through single-qubit Pauli operators (Fig. 1a). A pair of plaquette violations sharing an edge constitute a fermion, . We recently demonstrated the Abelian statistics of such quasiparticles in the surface code43. To realize non-Abelian statistics, one needs to go beyond such plaquette violations; it has been proposed that dislocations in the stabilizer graphanalogous to lattice defects in crystalline solidscan host projective non-Abelian Ising anyons9,10. For brevity, we refer to these as non-Abelian anyons or simply anyons from here on.

a, Stabilizer codes are conveniently described in a graph framework. Through deformations of the surface code graph, a square grid of qubits (crosses) can be used to realize more generalized graphs. Plaquette violations (red) correspond to stabilizers with sp=1 and are created by local Pauli operations. In the absence of deformations, plaquette violations are constrained to move on one of the two sublattices of the dual graph in the surface code, hence the two shades of blue. b, A pair of D3Vs (yellow triangles) appears by removing an edge between two neighbouring stabilizers, ({hat{S}}_{1}) and ({hat{S}}_{2}), and introducing the new stabilizer, (hat{S}={hat{S}}_{1}{hat{S}}_{2}). A D3V is moved by applying a two-qubit entangling gate, (exp left(frac{pi }{8}[{hat{S}}^{{prime} },hat{S}]right)). In the presence of bulk D3Vs, there is no consistent way of chequerboard colouring, hence the (arbitrarily chosen) grey regions. The top right shows that in a general stabilizer graph, ({hat{S}}_{p}) can be found from a constraint at each vertex, where {1,2}=0.

In the graph framework introduced above, it has been shown that such dislocations are characterized as vertices of degree 2 and 3 (ref. 23). Consider the stabilizer graph of the surface code26,44, specifically with boundary conditions such that the degeneracy is two. Although all the vertices in the bulk are D4Vs, one can create two D3Vs by removing an edge between two neighbouring plaquettes p and q, and introducing the new stabilizer (hat{S}={hat{S}}_{p}{hat{S}}_{q}) (Fig. 1b). Evidently, the introduction of two D3Vs reduces the number of independent stabilizers by one and thus doubles the degeneracy. This doubling is exactly what is expected when a pair of Ising anyons is introduced9,10; hence, D3Vs appear as a candidate of non-Abelian anyons, and we will denote them as .

To be braided and fused by unitary operations, the D3Vs must be moved. Whereas the structure of the stabilizer graph is usually considered to be static, it was predicted by Bombin that the dislocations in the surface code would show projective non-Abelian Ising statistics if braided10. Here, we will use a specific protocol recently proposed by Lensky et al.23 for deforming the stabilizer graph (and thus moving the anyons) using local two-qubit Clifford gates. To shift a D3V from vertex u to v, an edge must be disconnected from v and reconnected to u. This can be achieved by means of the gate unitary (exp left(frac{pi }{8}[{hat{S}}_{p}^{{prime} },{hat{S}}_{p}]right)), where ({hat{S}}_{p}) is the original stabilizer containing the edge and u, and ({hat{S}}_{p}^{{prime} }) is the new stabilizer that emerges after moving the edge23. In cases where the D3V is shifted between two connected vertices, the unitary simplifies to the form ({U}_{pm }({hat{tau }}_{u}{hat{tau }}_{v})equiv exp left(pm ifrac{pi }{4}{hat{tau }}_{u}{hat{tau }}_{v}right)), where ({hat{tau }}_{u}) and ({hat{tau }}_{v}) are Pauli operators acting on vertices u and v. We experimentally realize this unitary through a controlled Z (CZ) gate and single-qubit rotations (median errors of 7.3103 and 1.3103, respectively; Methods).

Following these insights from Kitaev and Bombin, we now turn to our experimental study of the proposed anyons, using the protocol described in ref. 23. In the first experiment, we demonstrate the creation of anyons and the fundamental fusion rules of and (Fig. 2a). In a 55 grid of superconducting qubits, we first use a protocol consisting of four layers of CZ gates to prepare the surface code ground state (Fig. 2b(i), see also ref. 43). The average stabilizer value after the ground state preparation is 0.940.04 (individual stabilizer values shown in Extended Data Fig. 3c). We then remove a stabilizer edge to create a pair of D3Vs () and separate them through the application of two-qubit gates. Fig. 2b(i)(iv) show the measured stabilizer values in the resultant graph in each step of this procedure (determined by simultaneously measuring the involved qubits in their respective bases, n=10,000; note that the measurements are destructive and the protocol is restarted after each measurement). In Fig. 2b(v), single-qubit Z gates are applied to two qubits near the lower left corner of the grid to create adjacent plaquette violations, which together form a fermion. Through the sequential application of X and Z gates (Fig. 2b(vii)(viii)), one of the plaquette violations is then made to encircle the right vertex. Crucially, after moving around , the plaquette violation does not return to where it started, but rather to the location of the other plaquette violation. This enables them to annihilate (Fig. 2b(viii)), causing the fermion to seemingly disappear. However, by bringing the two back together and annihilating them (Fig. 2b(ix)(xi)), we arrive at a striking observation: an particle re-emerges on two of the square plaquettes where the vertices previously resided.

a, The braiding worldlines used to fuse and . b, Expectation values of stabilizers at each step of the unitary operation after readout correction (see Extended Data Fig. 3 for details and individual stabilizer values). We first prepare the ground state of the surface code (step (i); average stabilizer value of 0.940.04, where the uncertaintyis one standard deviation). A D3V () pair is then created (ii) and separated (iii)(iv), before creating a fermion, (v). One of the plaquette violations is brought around the right (vi)(viii), allowing it to annihilate with the other plaquette violation (viii). The fermion has seemingly disappeared, but re-emerges when the are annihilated ((xi); stabilizer values 0.86 and 0.87). The path (v)(viii) demonstrates the fusion rule, =. The different fermion parities at the end of the paths (viii)(xi) and (iv)(i) show the other fusion rule, (sigma times sigma ={mathbb{1}}+varepsilon ). Yellow triangles represent the positions of the . The brown and red lines denote the paths of the and the plaquette violation, respectively. Red squares (diamonds) represent X (Z) gates. Upper left shows a table of two-qubit unitaries used in the protocol. Each stabilizer was measured n = 10,000 times in each step.c, A non-local technique for hidden fermion detection: the presence of a fermion in a -pair can be deduced by measuring the sign of the Pauli string (hat{{mathcal{P}}}) corresponding to bringing a plaquette violation around the -pair (grey path). (hat{{mathcal{P}}}) is equivalent to the shorter string (hat{{{mathcal{P}}}^{{prime} }}) (black path). Measurements of (hat{{{mathcal{P}}}^{{prime} }}) in steps (viii) (top) and (iv) (bottom) give values of 0.850.01 and +0.840.01, respectively. This indicates that there is a hidden fermion pair in the former case, but not in the latter, despite the stabilizers being the same.

Our results demonstrate the fusion of and . The disappearance of the fermion from step (v) to (viii) establishes the fundamental fusion rule of and :

$$sigma times varepsilon =sigma .$$

(2)

We emphasize that none of the single-qubit gates along the path of the plaquette violation are applied to the qubits hosting the mobile ; our observations are therefore solely due to the non-local effects of non-Abelian D3Vs, and exemplify the unconventional behaviour of the latter. Moreover, another fusion rule is seen by considering the reverse path (iv)(i), and comparing it to the path (viii)(xi). These two paths demonstrate that a pair of can fuse to form either vacuum (({mathbb{1}})) or one fermion (steps (i) and (xi), respectively):

$$sigma times sigma ={mathbb{1}}+varepsilon .$$

(3)

The starting points of these two paths ((iv) and (viii)) cannot be distinguished by any local measurement. We therefore introduce a non-local measurement technique that allows for detecting an without fusing the (refs. 10,23,26). The key idea underlying this method is that bringing a plaquette violation around a fermion should result in a phase. We therefore measure the Pauli string (hat{{mathcal{P}}}) that corresponds to creating two plaquette violations, bringing one of them around the two , and finally annihilating them with each other (grey paths in Fig. 2c). The existence of an inside the -pair should cause (langle hat{{mathcal{P}}}rangle =-1). To simplify this technique further, (hat{{mathcal{P}}}) can be reduced to a shorter string (hat{{{mathcal{P}}}^{{prime} }}) (black paths in Fig. 2c) by taking advantage of the stabilizers it encompasses. For instance, if (hat{{mathcal{P}}}) contains three of the operators in a four-qubit stabilizer, these can be switched out with the remaining operator. Measuring (langle hat{{{mathcal{P}}}^{{prime} }}rangle ) instep (iv), in which the are separated but the fermion has not yet been introduced, gives (langle hat{{{mathcal{P}}}^{{prime} }}rangle =+,0.84pm 0.01), consistent with the absence of fermions(Fig. 2c). However, performing the exact same measurement in step(viii), in which the are in the same positions, we find (langle hat{{{mathcal{P}}}^{{prime} }}rangle =-,0.85pm 0.01), indicating that an is delocalized across the spatially separated pair(Fig. 2c). This observation highlights the non-local encoding of the fermions, which cannot be explained with classical physics.

Having demonstrated the above fusion rules involving , we next braid them with each other to directly show their non-Abelian statistics. We consider two spatially separated pairs, A and B, by removing two stabilizer edges (Fig. 3a,b(ii)). Next, we apply two-qubit gates along a horizontal path to separate the in pair A (Fig. 3b(iii)), followed by a similar procedure in the vertical direction on pair B (Fig. 3b(iv)), so that one of its crosses the path of pair A. We then subsequently bring the from pairs A and B back to their original positions (Fig. 3b(v)(viii) and (ix)(xi), respectively). When the two pairs are annihilated in the final step (Fig. 3b(xii)), we observe that a fermion is revealed in each of the positions where the pairs resided (average stabilizer value 0.450.06). This shows a clear change in local observables from the initial state in which no fermions were present. As a control experiment, we repeat the experiment with distinguishable pairs, achieved by attaching a plaquette violation to each of the in pair B (Fig. 3c,d; see also Extended Data Fig. 8 for stabilizer measurements through the full protocol). Moving the plaquette violation along with the requires a string of single-qubit gates, which switches the direction of the rotation in the multiqubit unitaries, UU. In this case, no fermions are observed at the end of the protocol (average stabilizer value +0.460.04), thus providing a successful control.

a, Wordline schematic of the braiding process. b, Experimental demonstration of braiding, showing the values of the stabilizers throughout the process. Two pairs, A and B, are created from the vacuum ({mathbb{1}}), and one of the in pair A is brought to the right side of the grid. Next, a from pair B is moved to the top, thus crossing the path of pair A, before bringing pairs A and B back together to complete the braid. In the final step, two fermions appear in the locations where the pairs resided, constituting a change in the local observables. The diagonal move in step (iv) requires two SWAP gates (three CZ gates each) and a total of ten CZ gates. The three-qubit unitary in step (viii) requires four SWAP gates and a total of 15 CZ gates. In the full circuit, a total of 40 layers of CZ gates are applied (Methods). The yellow triangles represent the locations of the ; the brown and green lines represent the paths of from pairs A and B, respectively.The four red stabilizers in (xii) have a mean value of 0.45 0.06, where the uncertaintyis one standard deviation. Each stabilizer was measured n= 10,000 timesin each step.c, As a control experiment, we perform the same braid as in a, but with distinguishable by attaching a plaquette violation to the in pair B (represented with purple triangles). d, Same as b, but using distinguishable (only showing steps (i), (iv) and (xii)). In contrast to b, no fermions are observed in step (xii).

Fermions can only be created in pairs in the bulk. Moreover, the fusion of two can only create zero or one fermion (equation (3)). Hence, our experiment involves the minimal number of bulk (four) needed to encode two fermions and demonstrate non-Abelian braiding. Because the fermion parity is conserved, effects of gate imperfections and decoherence can be partially mitigated by postselecting for an even number of fermions. This results in fermion detection values of 0.760.03 and +0.790.04 in Fig. 3b,d, respectively.

Together, our observations show the change in local observables by braiding of indistinguishable and constitute a direct demonstration of their non-Abelian statistics. In other words, the double-braiding operation R2 is a matrix that cannot be reduced to a scalar. Specifically, it corresponds to an X gate acting on the space spanned by zero- and two-fermion wavefunctions.

The full braiding circuit consists of 40 layers of CZ gates and 41 layers of single-qubit gates (36 of each after ground state preparation). The effects of imperfections in this hardware implementation can be assessed through comparison with the control experiment. The absolute values of the stabilizers in which the fermions are detected in the two experiments (dashed boxes in Fig. 3b,d(xii)) are very similar (average values of 0.45 and +0.46). This is consistent with the depolarization channel model, in which the measured stabilizer values are proportional to the ideal values (1).

We next study the prospects of using D3Vs to encode logical qubits and prepare an entangled state of anyon pairs. By doubling the degeneracy, each additional pair introduces one logical qubit, where the ({left|0rightrangle }_{{rm{L}}}) (({| 1rangle }_{{rm{L}}})) state corresponds to an even (odd) number of hidden fermions. The measurements of the fermion numbers in several pairs are not fully independent: bringing a plaquette violation around one pair is equivalent to bringing it around all the other pairs (due to the conservation of fermionic parity). Hence, N2 anyons encode N/21 logical qubits. The D3Vs we have created and manipulated so far are not the only ones present in the stabilizer graph; with the boundary conditions used here, each of the four corners are also D3Vs, no different from those in the bulk23. Indeed, the existence of D3Vs in the corners is the reason why a single fermion could be created in the corner in Fig. 2b(v). This is also consistent with the fact that the surface code itself encodes one logical qubit in the absence of additional D3Vs. Here we create two pairs of D3Vs, in addition to the four that are already present in the corners, to encode a total of three logical qubits.

Through the use of braiding, we aim to prepare an entangled state of these logical qubits, specifically a GHZ state on the form ((| 000rangle +| 111rangle )/sqrt{2}). The definition of a GHZ state and the specifics of how it is prepared is basis-dependent. In most systems, the degrees of freedom are local and there is a natural choice of basis. For spatially separated anyons, the measurement operators are necessarily non-local. Here we choose the basis defined as follows: for the first two logical qubits, we choose the logical ({hat{Z}}_{{rm{L}},i}) operators to be Pauli strings encircling each of the bulk pairs, as was used in Fig. 2c (green and turquoise paths in the left column of Fig. 4a). For the logical surface code qubit, we define ({hat{Z}}_{{rm{L}},3}) as the Pauli string that crosses the grid horizontally through the gap between the bulk D3V pairs, effectively enclosing four (red path in Fig. 4a). In this basis, the initial state is a product state.

a, Logical operators of the three logical qubits encoded in the eight anyons (other basis choices are possible). The coloured curves in the left column denote plaquette violation paths, before reduction to shorter, equivalent Pauli strings measured in the experiment (right column). b, Worldline schematic of the single exchange used to realize an entangled state of the logical qubits. c, Single exchange of the non-Abelian anyons, showing measurements of the stabilizers throughout the protocol. Yellow triangles represent the locations of the , whereas brown and green lines denote their paths. The average stabilizer values are 0.950.04 and 0.880.05 (one standard deviation)in the first and last step, respectively.Each stabilizer was measuredn= 20,000 timesin each step. d,e, Real (d) and imaginary (e) parts of the reconstructed density matrix from the quantum state tomography. ({rm{Re}}(rho )) has clear peaks in its corners, as expected for a GHZ state on the form ((left|000rightrangle +left|111rightrangle )/sqrt{2}). The overlap with the ideal GHZ state is ({rm{Tr}}{,{rho }_{{rm{GHZ}}}rho }=0.623pm 0.004), where the uncertaintyis one standard deviation determined from bootstrapping.

Whereas a double braid was used to implement the operator X in Fig. 3, we now perform a single braid (Fig. 4b) to realize (sqrt{X}) and create a GHZ state. We implement this protocol by bringing one from each bulk pair across the grid to the other side (Fig. 4c). For every anyon double exchange across a Pauli string, the value of the Pauli string changes sign. Hence, a double exchange would change (left|000rightrangle ) to (left|111rightrangle ), whereas a single exchange is expected to realize the superposition, ((left|111rightrangle +left|000rightrangle )/sqrt{2}).

To study the effect of this operation, we perform quantum state tomography on the final state, which requires measurements of not only ({hat{Z}}_{{rm{L}},i}), but also ({hat{X}}_{{rm{L}},i}) and ({hat{Y}}_{{rm{L}},i}) on the three logical qubits. For the first two logical qubits, ({hat{X}}_{{rm{L}},i}) is the Pauli string that corresponds to bringing a plaquette violation around only one of the in the pair (as demonstrated in Fig. 2b). Both the logical ({hat{X}}_{{rm{L}},i}) and ({hat{Z}}_{{rm{L}},i}) operators can be simplified by reducing the original Pauli strings (green and turquoise paths in the left column of Fig. 4c) to equivalent, shorter ones (right column). ({hat{Z}}_{{rm{L}},1}) can in fact be reduced to a single (hat{Y})-operator. For the logical surface code qubit, we define ({hat{X}}_{{rm{L}},3}) as the Pauli string that crosses the grid vertically between the bulk D3V pairs (red path in Fig. 4a). Finally, the logical ({hat{Y}}_{{rm{L}},i})-operators are simply found from ({hat{Y}}_{{rm{L}},i}=i{hat{X}}_{{rm{L}},i}{hat{Z}}_{{rm{L}},i}). Measuring these operators, we reconstruct the density matrix of the final state (Fig. 4d,e), which has a purity of (sqrt{{rm{Tr}}{{rho }^{2}}}=0.646pm 0.003) and an overlap with the ideal GHZ state of ({rm{Tr}}{{rho }_{{rm{GHZ}}}rho }=0.623pm 0.004) (uncertainties estimated from bootstrapping method; resampled 10,000 times from the original data set). The fact that the state fidelity is similar to the purity suggests that the infidelity is well described by a depolarizing error channel.

In conclusion, we have realized highly controllable braiding of degree-3 vertices, enabling the demonstration of the fusion and braiding rules of non-Abelian Ising anyons. We have also shown that braiding can be used to create an entangled state of three logical qubits encoded in these anyons. In other, more conventional candidate platforms for non-Abelian exchange statistics, which involve Hamiltonian dynamics of quasi-particle excitations, topological protection naturally arises from an emergent gap that separates the computational states from other states. To leverage the non-Abelian anyons in our system for topologically protected quantum computing, the stabilizers must be measured throughout the braiding protocol. The potential inclusion of this error correction procedure, which involves overheads including readout of five-qubit stabilizers, could open a new path towards fault-tolerant implementation of Clifford gates, a key ingredient of universal quantum computation.

Read the original:
Non-Abelian braiding of graph vertices in a superconducting processor - Nature.com

Google Quantum AI Breaks Ground: Unraveling the Mystery of Non-Abelian Anyons – Neuroscience News

Summary: For the first time, Google Quantum AI has observed the peculiar behavior of non-Abelian anyons, particles with the potential to revolutionize quantum computing by making operations more resistant to noise.

Non-Abelian anyons have the unique feature of retaining a sort of memory, allowing us to determine when they have been exchanged, even though they are identical.

The team successfully used these anyons to perform quantum computations, opening a new path towards topological quantum computation. This significant discovery could be instrumental in the future of fault-tolerant topological quantum computing.

Key Facts:

Source: Google Quantum AI

Our intuition tells us that it should be impossible to see whether two identical objects have been swapped back and forth, and for all particles observed to date, that has been the case. Until now.

Non-Abelian anyons the only particles that have been predicted to break this rule have been sought for their fascinating features and their potential to revolutionize quantum computing by making the operations more robust to noise.

Microsoft and others have chosen this approach for their quantum computing effort. But after decades of efforts by researchers in the field, observing non-Abelian anyons and their strange behavior has proven challenging, to say the least.

In apaperposted on the preprint server Arxiv.org last October andpublishedinNaturetoday, researchers at Google Quantum AI announced that they had used one of their superconducting quantum processors to observe the peculiar behavior of non-Abelian anyons for the first time ever.

They also demonstrated how this phenomenon could be used to perform quantum computations. Earlier this week the quantum computing company Quantinuum released another study on the topic, complementing Googles initial discovery.

These new results open a new path toward topological quantum computation, in which operations are achieved by winding non-Abelian anyons around each other like strings in a braid.

Google Quantum AI team member and first author of the manuscript, Trond I. Andersen says, Observing the bizarre behavior of non-Abelian anyons for the first time really highlights the type of exciting phenomena we can now access with quantum computers.

Imagine youre shown two identical objects and then asked to close your eyes. Open them again, and you see the same two objects. How can you determine if they have been swapped? Intuition says that if the objects are truly identical, there is no way to tell.

Quantum mechanics supports this intuition, but only in our familiar three-dimensional world. If the identical objects are restricted to only move in a two-dimensional plane, sometimes, our intuition can fail and quantum mechanics allows for something bizarre: non-Abelian anyons retain a sort of memory it is possible to tell when two of them have been exchanged, despite being completely identical.

This memory of the non-Abelian anyons can be thought of as a continuous line in space-time: the particles so-called world-line. When two non-Abelian anyons are exchanged, their world-lines wrap around one another. Wrap them in the right way, and the resulting knots and braids form the basic operations of a topological quantum computer.

The team started by preparing their superconducting qubits in an entangled quantumstate that is well represented as a checkerboard a familiar configuration for the Google team, who recentlydemonstrated a milestone in quantum error correctionusing this setup. In the checkerboard arrangement, related but less useful particles called Abelian anyons can emerge.

To realize non-Abelian anyons, the researchers stretched and squashed the quantum state of their qubits to transform the checkered pattern into oddly shaped polygons. Particular vertices in these polygons hosted the non-Abelian anyons.

Using aprotocoldeveloped by Eun-Ah Kim at Cornell University and former postdoc Yuri Lensky, the team could then move the non-Abelian anyons around by continuing to deform the lattice and shifting the locations of the non-Abelian vertices.

In a series of experiments, the researchers at Google observed the behavior of these non-Abelian anyons and how they interacted with the more mundane Abelian anyons.

Weaving the two types of particles around one another yielded bizarre phenomena particles mysteriously disappeared, reappeared and shapeshifted from one type to another as they wound around one another and collided.

Most importantly, the team observed the hallmark of non-Abelian anyons: when two of them were swapped, it caused a measurable change in the quantum state of their system a striking phenomenon that had never been observed before.

Finally, the team demonstrated how braiding of non-Abelian anyons might be used in quantum computations. By braiding several non-Abelian anyons together, they were able to create a well-known quantum entangled state called the Greenberger-Horne-Zeilinger (GHZ) state.

The physics of non-Abelian particles is also at the core of the approach that Microsoft has chosen for their quantum computing effort. While they are attempting to engineer material systems that intrinsically host these anyons, the Google team has now shown that the same type of physics can be realized on their superconducting processors.

This week the quantum computing company Quantinuum released an impressive complementary study that also demonstrated non-Abelian braiding, in this case using a trapped-ion quantum processor. Andersen is excited to see other quantum computing groups observing non-Abelian braiding as well.

He says, It will be very interesting to see how non-Abelian anyons are employed in quantum computing in the future, and whether their peculiar behavior can hold the key to fault-tolerant topological quantum computing.

Author: Katie McCormickSource: Google Quantum AIContact: Katie McCormick Google Quantum AIImage: The image is credited to Neuroscience News

Original Research: Open access.Non-Abelian braiding of graph vertices in a superconducting processor by Trond I. Andersen et al. Nature

Abstract

Non-Abelian braiding of graph vertices in a superconducting processor

Indistinguishability of particles is a fundamental principle of quantum mechanics. For all elementary and quasiparticles observed to dateincluding fermions, bosons and Abelian anyonsthis principle guarantees that the braiding of identical particles leaves the system unchanged.

However, in two spatial dimensions, an intriguing possibility exists: braiding of non-Abelian anyons causes rotations in a space of topologically degenerate wavefunctions. Hence, it can change the observables of the system without violating the principle of indistinguishability.

Despite the well-developed mathematical description of non-Abelian anyons and numerous theoretical proposals, the experimental observation of their exchange statistics has remained elusive for decades. Controllable many-body quantum states generated on quantum processors offer another path for exploring these fundamental phenomena.

Whereas efforts on conventional solid-state platforms typically involve Hamiltonian dynamics of quasiparticles, superconducting quantum processors allow for directly manipulating the many-body wavefunction by means of unitary gates.

Building on predictions that stabilizer codes can host projective non-Abelian Ising anyons, we implement a generalized stabilizer code and unitary protocolto create and braid them.

This allows us to experimentally verify the fusion rules of the anyons and braid them to realize their statistics. We then study the prospect of using the anyons for quantum computation and use braiding to create an entangled state of anyons encoding three logical qubits.

Our work provides new insights about non-Abelian braiding and, through the future inclusion of error correction to achieve topological protection, could open a path towards fault-tolerant quantum computing.

Go here to see the original:
Google Quantum AI Breaks Ground: Unraveling the Mystery of Non-Abelian Anyons - Neuroscience News

IonQ Announces Participation in 18th Annual Needham Technology & Media Conference – Yahoo Finance

COLLEGE PARK, Md., May 12, 2023--(BUSINESS WIRE)--IonQ, Inc. (NYSE: IONQ), a leader in quantum computing, today announced that Thomas Kramer, Chief Financial Officer, and Jordan Shapiro, Vice President of FP&A and Head of Investor Relations, will participate in a fireside chat at the 18th Annual Needham Technology & Media Conference at the Intercontinental New York Times Square in New York City on Tuesday, May 16, 2023. The Companys discussion will begin at 10:15 AM ET and the webcast link will be available on our Companys website here, or directly here.

About IonQIonQ, Inc. is a leader in quantum computing, with a proven track record of innovation and deployment. IonQs current generation quantum computer, IonQ Forte, is the latest in a line of cutting-edge systems, boasting an industry-leading 29 algorithmic qubits. Along with record performance, IonQ has defined what it believes is the best path forward to scale.

IonQ is the only company with its quantum systems available through the cloud on Amazon Braket, Microsoft Azure and Google Cloud, as well as through direct API access. IonQ was founded in 2015 by Christopher Monroe and Jungsang Kim based on 25 years of pioneering research. To learn more, visit http://www.ionq.com.

View source version on businesswire.com: https://www.businesswire.com/news/home/20230512005006/en/

Contacts

IonQ Media:press@ionq.co

IonQ Investor:investors@ionq.co

Read more:
IonQ Announces Participation in 18th Annual Needham Technology & Media Conference - Yahoo Finance