Fact-checked by Grok 2 weeks ago

Phase qubit

A phase qubit is a type of superconducting quantum bit (qubit) implemented using a current-biased Josephson junction shunted by a large on-chip capacitance, forming a nonlinear LC oscillator where the phase difference across the junction acts as the primary degree of freedom. The qubit states are encoded in the two lowest energy eigenstates of the anharmonic potential well created by the Josephson junction, with transitions driven by microwave pulses in the 4–8 GHz range. Operated at millikelvin temperatures to maintain superconductivity, phase qubits exhibit macroscopic quantum behavior, including coherent oscillations and entanglement, making them suitable for early quantum information processing demonstrations. The concept of the phase qubit originated from early studies of quantum tunneling in Josephson junctions in the 1980s, but was formally proposed as a scalable design in 2001 by researchers at the University of Maryland, who suggested encoding in the ground and first excited states of a current-biased junction. The first experimental demonstration came in 2002 at the National Institute of Standards and Technology (NIST), where Rabi oscillations—coherent flips between qubit states—were observed in a large-area Josephson junction device, confirming its quantum nature with oscillation frequencies up to 70 MHz and coherence times on the order of nanoseconds. Subsequent advancements, particularly at the (UCSB), refined fabrication using aluminum-based junctions on substrates, enabling integration with circuitry. Key characteristics of phase qubits include their of approximately 50 Ω, which matches microwave transmission lines for efficient readout and , and their large shunt (typically 100–500 pF), which allows strong dispersive to resonators with coupling strengths g/2π exceeding 100 MHz. These features support multi-qubit operations, such as the 2006 demonstration of entanglement between two phase qubits via , achieving fidelities above 70%. By 2008, experiments had generated nonclassical Fock states with up to 15 photons in qubit-resonator systems and improved times to T₁ ≈ 600 ns for energy relaxation and T₂ ≈ 200 ns for , limited primarily by losses in the . Phase qubits played a pivotal role in early research, paving the way for more advanced designs. Although phase qubits demonstrated key quantum phenomena, they have been largely supplanted by more advanced superconducting qubit designs, such as the , which exhibit longer times and greater .

Introduction

Definition and Operating Principle

A phase qubit is a type of superconducting that encodes in the phase difference across a current-biased Josephson junction operating in the zero-voltage state. The device consists of a single Josephson junction shunted by a , forming a nonlinear where the variable serves as the primary degree of freedom for quantum operations. The operating principle relies on the anharmonic potential created by the Josephson junction, which resembles a tilted washboard landscape. The two lowest-energy metastable states within a potential well are used as the computational basis states |0⟩ (ground state) and |1⟩ (first excited state). Superposition and entanglement emerge from quantum tunneling between these states, enabled by microwave pulses that drive coherent transitions while the anharmonicity prevents unwanted higher-level involvement. A bias current tunes the energy landscape by tilting the potential wells, adjusting the barrier height and thus controlling the qubit's escape rates for state preparation and readout. Here, the phase difference φ across acts as the conjugate degree of freedom, analogous to the of a particle in a quantum potential model. To maintain and suppress thermal excitations, phase qubits require operation at cryogenic temperatures near 20 mK in dilution refrigerators.

Historical Development

The development of the phase qubit emerged in the early within the broader context of superconducting qubit research, building on the groundbreaking 1999 demonstration of coherent quantum control in a charge qubit by Nakamura, Pashkin, and Tsai using a box. This work highlighted the potential of Josephson junctions for macroscopic , inspiring parallel explorations into phase-based and flux-based designs to address limitations like charge noise sensitivity. The phase qubit, which encodes information in the phase difference across a current-biased Josephson junction, represented an effort to leverage the nonlinear of the junction for qubit operations while operating near an escape point in the phase potential. The first experimental realization of the phase qubit was achieved in 2002 by and colleagues at NIST, who demonstrated coherent Rabi oscillations in a large Josephson junction , along with initialization and single-shot readout via macroscopic quantum tunneling. This milestone confirmed the viability of phase states for processing, with initial coherence times around 20 , limited primarily by and in the bias circuitry. Influenced by contemporaneous advances in charge and flux qubits, such as the 2002 quantronium demonstration by Devoret's group, the phase qubit quickly progressed to multi-qubit interactions; by 2006, Steffen et al. reported the creation and tomographic verification of entangled Bell states between two capacitively coupled phase qubits, enabling two-qubit gates with fidelities up to 87%. Further refinements in the mid-2000s focused on mitigating decoherence sources, notably dielectric losses from two-level systems (TLS) in amorphous materials, as identified by Martinis et al. in 2005, leading to improved energy relaxation times (T₁) exceeding 500 through optimized like a-Si:H. Key contributions came from teams led by John Martinis (initially at NIST, later UC Santa Barbara), Ken Segall (Yale/ collaborations on related phase dynamics), and groups at Yale under Robert Schoelkopf. Readout techniques also advanced around 2007–2008, incorporating faster switching SQUIDs and reduced crosstalk for higher-fidelity measurements, as explored in process experiments. By the 2010s, phase qubits were largely superseded by the design, proposed by Koch et al. in , which offered superior charge noise insensitivity and scalability for larger arrays due to its higher and relaxed fabrication tolerances. While phase qubits achieved coherence times up to microseconds in optimized setups and informed hybrid qubit architectures combining phase and flux elements, their sensitivity to flux noise and challenges in multi-qubit coupling limited widespread adoption in fault-tolerant efforts.

Physical Foundations

Superconductor-Insulator-Superconductor Junction

The superconductor-insulator-superconductor () junction forms the core weak link in qubits, comprising two superconducting electrodes separated by a thin insulating barrier, typically aluminum oxide (AlO_x) with a thickness of 1-2 nm, which enables the tunneling of pairs across the barrier while maintaining coherence between the superconductors. Fabrication of these junctions for phase qubits involves electron-beam lithography to pattern aluminum films, followed by controlled thermal or plasma oxidation to grow the insulating AlO_x layer, yielding high critical current densities J_c \sim 0.1{-}0.2\,\mu\mathrm{A}/\mu\mathrm{m}^2 that support the high-frequency plasma oscillations necessary for qubit operation. Classically, the SIS junction obeys the DC Josephson relation, which connects the voltage V across the junction to the time derivative of the superconducting phase difference \phi as V = \frac{\hbar}{2e} \frac{d\phi}{dt}, and the AC Josephson relation, which governs the supercurrent I as I = I_c \sin(\phi), where I_c is the critical current, \hbar is the reduced Planck's constant, and e is the elementary charge. The voltage-current characteristics of an SIS junction feature a zero-voltage supercurrent branch for currents below I_c, enabling dissipationless flow; above I_c, the junction enters a resistive state where voltage develops due to tunneling through the , often manifesting as abrupt voltage jumps in underdamped configurations. For phase qubits, SIS junctions are engineered to be underdamped, exhibiting a quality factor Q > 10^4, which reduces energy loss and promotes the macroscopic quantum tunneling essential to the qubit's phase dynamics under bias.

Josephson Effect and Phase Dynamics

The Josephson effect describes the quantum mechanical tunneling of Cooper pairs across a thin insulating barrier in a superconductor-insulator-superconductor (SIS) junction, resulting in a dissipationless supercurrent that flows without an applied voltage. This supercurrent is expressed as I = I_c \sin \phi, where I_c is the critical current and \phi is the gauge-invariant phase difference between the superconducting wave functions on either side of the junction. The temporal evolution of the phase follows the second Josephson relation, \frac{d\phi}{dt} = \frac{2eV}{\hbar}, linking the rate of phase change to the voltage V across the junction, with e the elementary charge and \hbar the reduced Planck's constant. These relations, derived from a quantum tunneling model, enable coherent phase dynamics essential to superconducting devices. In a phase qubit, the SIS junction is biased by a I_b such that I_b < I_c, which introduces a linear tilt to the otherwise periodic potential energy landscape governing the phase \phi. This tilted washboard potential is given by U(\phi) = -E_J \left[ \cos \phi + \left( \frac{I_b}{I_c} \right) \phi \right], where E_J = \frac{\hbar I_c}{2e} represents the Josephson energy scale, setting the depth of the potential wells. The bias current modulates the tilt angle, creating a series of metastable minima separated by barriers; for I_b approaching I_c, the barriers diminish, allowing controlled manipulation of the phase particle's position within a well. This configuration underpins the classical dynamics of the phase qubit, where the phase acts as the primary degree of freedom analogous to a particle in a biased periodic potential. The dynamics of the phase in the zero-voltage state involve small oscillations around a potential minimum, characterized by the plasma frequency \omega_p = \sqrt{\frac{2e I_c}{\hbar C}}, with C the junction capacitance; this frequency dictates the natural oscillation period of the phase particle in the cubic approximation to the well. Escape from these metastable states occurs through phase slips, where the phase advances by $2\pi, transitioning to the voltage state; such slips can be thermally activated at higher temperatures or, at low temperatures, proceed via quantum processes. The minima correspond to circulating supercurrents in the junction loop, with the barrier height between adjacent wells approximated as \Delta U \approx \frac{2 E_J}{\pi} \left(1 - \frac{I_b}{I_c}\right)^{3/2} near I_b \approx I_c, providing a tunable energy scale for state control. A critical aspect of phase dynamics is macroscopic quantum tunneling (MQT) of the collective phase variable, where the phase particle tunnels through the inverted parabolic barrier rather than surmounting it classically, even at temperatures near absolute zero. Observed in underdamped junctions, MQT rates align with quantum mechanical predictions incorporating dissipation, enabling the phase qubit's operation by facilitating transitions between low-lying energy states within the well. This phenomenon, first experimentally verified in current-biased junctions in the 1980s by groups including those of , , and —who shared the 2025 Nobel Prize in Physics for these demonstrations—highlights the quantum coherence of macroscopic superconducting systems and serves as the basis for qubit state manipulation.

Circuit Modeling

Resistively and Capacitively Shunted Junction Model

The Resistively and Capacitively Shunted Junction (RCSJ) model describes the classical dynamics of a Josephson junction by treating it as an ideal nonlinear inductor shunted in parallel by a linear resistor R and capacitor C. The resistor models quasiparticle dissipation and damping, while the capacitor represents the junction's geometric capacitance, introducing inertial effects from charge accumulation. In the phase qubit configuration, the equivalent circuit features a current-biased shunted , where the bias current I_b flows through the parallel combination. This current divides into the supercurrent I_c \sin \phi across the ideal junction, the ohmic current V/R through the shunt resistor, and the displacement current C \, dV/dt through the capacitor, with the voltage V = (\hbar / 2e) \, d\phi / dt relating the phase difference \phi to the time-varying voltage via the second . The resulting dynamics are governed by a second-order differential equation analogous to a damped, driven particle in a washboard potential: \frac{\hbar C}{2e} \frac{d^2 \phi}{dt^2} + \frac{\hbar}{2e R} \frac{d\phi}{dt} + I_c \sin \phi = I_b. This equation captures the phase particle's motion, with the inertial term from capacitance, damping from resistance, and nonlinear restoring force from the Josephson inductance. For phase qubits, the RCSJ model highlights the importance of shunting to achieve an underdamped regime (\beta_c \gg 1), which reduces quasiparticle damping and enables macroscopic quantum tunneling and coherent plasma oscillations at the qubit's escape frequency. This underdamping is realized in phase qubits using large-area junctions where the Josephson energy E_J exceeds the charging energy E_C by 10 to 100, yielding plasma frequencies in the 5–10 GHz range suitable for microwave control and readout. The RCSJ model originated in the late 1960s to explain hysteretic current-voltage characteristics in early Josephson junctions and was later adapted in the early 2000s for analyzing the macroscopic quantum behavior in superconducting phase qubits.

McCumber-Stewart Parameters

The Stewart-McCumber parameters are key dimensionless quantities arising from the RCSJ model that quantify the damping and hysteretic behavior of the Josephson junction in phase qubits. These parameters govern the classical dynamics of the phase across the junction, influencing the stability of metastable states essential for qubit operation. The central parameter is the Stewart-McCumber parameter \beta_c = \frac{2\pi I_c R^2 C}{\Phi_0}, where \Phi_0 = h/2e is the magnetic flux quantum, I_c is the critical current of the junction, R is the shunt (or normal-state) resistance, and C is the total junction capacitance. This parameter characterizes the degree of underdamping in the junction: values of \beta_c > 1 indicate hysteretic behavior in the current-voltage (I-V) characteristics, allowing the phase to switch between voltage and zero-voltage states with a nonzero return current. In phase qubits, \beta_c \gg 1 is required to create deep potential wells in the tilted washboard potential, enabling the phase particle to oscillate coherently between discrete energy levels without rapid thermal escape. From the RCSJ model, \beta_c directly determines the I_r in the hysteretic I-V , approximated as I_r \approx I_c \left(4/\beta_c\right)^{1/4} for large \beta_c, which sets the threshold for retrapping into the superconducting state after voltage switching. High \beta_c thus supports multiple metastable states near the top of the cubic , where the two lowest energy levels serve as the states |0\rangle and |1\rangle. In experimental phase implementations, typical junction parameters yield \beta_c \sim 10^3 to $10^5: I_c \sim 1--$10~\muA, R \sim 10--$100~\Omega, and C \sim 50--$100 pF. A related figure of merit is the quality factor Q = \omega_p R C, where \omega_p = \sqrt{2\pi I_c / (\Phi_0 C)} is the characteristic frequency of the unbiased . This parameter measures the damping relative to the frequency, with Q = \sqrt{\beta_c} providing a direct link to the Stewart-McCumber parameter. For phase qubits, Q \sim 10^4--$10^5 is targeted to sufficient times (on the of hundreds of nanoseconds) while allowing controllable microwave-driven transitions between qubit states. In single-junction phase qubits, the inductance parameter \beta_L = 2\pi L I_c / \Phi_0 (with L the loop ) plays a minimal role, as the design avoids closed superconducting loops to prevent -related dynamics; any incidental inductance is kept small such that \beta_L \ll 1. This contrasts with qubits, where \beta_L \approx 1 enables tunable .

Nonlinear Behavior

The Nonlinear Inductor

The Josephson junction functions as the primary nonlinear inductor in the phase qubit circuit, enabling anharmonic behavior that defines the qubit's computational states through phase dynamics across . The effective inductance of the junction is phase-dependent, expressed as
L_J(\phi) = \frac{\Phi_0}{2\pi I_c \cos \phi},
where \Phi_0 = h/(2e) is the , I_c is the critical current, and \phi is the superconducting phase difference. This variation with \phi imparts the essential nonlinearity, as the inductance decreases from its maximum at \phi = 0 and diverges near \phi = \pi/2.
The energy stored in this inductor corresponds to the magnetic energy
E_L = \frac{ (\Phi_0 \phi / 2\pi)^2 }{2 L_J },
which exhibits nonlinearity arising from the \cos \phi term in L_J, contrasting with the quadratic energy of a linear inductor.
For small phase excursions around \phi = 0, the small-signal inductance reduces to L_0 = \Phi_0 / (2\pi I_c), providing a baseline linear response. The inherent anharmonicity of L_J(\phi) ensures non-equally spaced energy levels when quantized, which is vital for isolating the two lowest states as the qubit manifold. Phase qubits employ large-area Josephson junctions to maximize E_J = I_c \Phi_0 / (2\pi), typically achieving a nonlinearity \alpha = E_J / E_C \sim 10^4, where E_C = e^2 / (2 C_J) is the charging of the junction C_J. This high enhances the delocalization of the variable, minimizing charge noise effects and stabilizing the as the degree of freedom. In contrast to flux qubits, which utilize linear inductors in closed superconducting loops to encode information in persistent currents, the phase-dependent nonlinearity of the Josephson inductor permits direct manipulation of the phase \phi as the qubit variable.

Tilted Washboard Potential

In the resistively and capacitively shunted junction (RCSJ) model, the effective potential energy for the superconducting phase difference φ across a current-biased Josephson junction is derived from the Josephson coupling energy combined with the work done by the bias current. This yields the tilted washboard potential U(\phi) = E_J \left[ -\cos \phi + s \phi \right], where E_J = \frac{\hbar I_c}{2e} is the Josephson energy scale, I_c is the critical current, and s = I_b / I_c (with $0 < s < 1) is the normalized bias current that introduces the linear tilt. The cosine term provides the periodic structure with period $2\pi, while the linear term s \phi causes the overall potential to decrease indefinitely with increasing φ, creating a series of metastable wells that decrease in depth along the tilt direction. This shape evokes an inclined washboard, where the phase φ acts as the position of a classical particle, and the zero-voltage state corresponds to the particle being trapped in one of these shallow wells without escaping to generate a voltage. The local minima of the potential, representing stable zero-voltage states, occur at phase values \phi_\min \approx \arcsin(s) + 2\pi n for integer n, while the intervening maxima are at \phi_\max = \pi - \arcsin(s) + 2\pi n. The barrier height separating adjacent minima and maxima, which determines the stability of the trapped state, is given by \Delta U = 2 E_J \left[ (1 - s^2)^{1/2} - s \arccos(s) \right]. This expression shows that \Delta U decreases monotonically as s increases toward 1, eventually vanishing at s = 1 when the bias reaches the and all wells disappear. Near the phase qubit's operating regime, where s \approx 0.8, the well approximates a cubic potential locally, enhancing anharmonicity for defining qubit states while keeping the barrier sufficiently high for coherence. Classically, escape from a minimum over the barrier is governed by thermal activation, with the rate approximated as \Gamma \approx \omega_p \exp(-\Delta U / k_B T), where \omega_p = \sqrt{2 e I_c / \hbar C} (1 - s^2)^{1/4} is the small-oscillation plasma frequency at the well bottom (with C the junction capacitance) and k_B T the thermal energy. At low temperatures where k_B T \ll \Delta U / \hbar \omega_p, thermal activation is suppressed, and escape proceeds via macroscopic quantum tunneling (MQT) through the barrier, a process first observed in and central to phase qubit readout. The MQT rate depends on the barrier shape and damping but follows a similar exponential form modulated by quantum prefactors. By tuning the bias current s, the barrier height \Delta U can be controllably reduced, allowing the escape rate to be adjusted over orders of magnitude for qubit manipulation and measurement; typical operation occurs at s \sim 0.7--$0.9 to balance sufficient barrier height for state storage (e.g., \Delta U \sim 5--$10 k_B T) with fast readout via escape. This regime exploits the zero-voltage metastable states in the shallowest wells of the periodic, downward-tilting landscape, where the phase particle oscillates harmonically at low energies but anharmonically at higher ones, enabling superposition of the two lowest states as the qubit basis.

Quantum Mechanics

Circuit Quantization

The quantization of the resistively and capacitively shunted Josephson junction (RCSJ) circuit, which forms the basis of the phase qubit, proceeds by treating the superconducting phase difference \phi across the junction as a canonical coordinate analogous to position in a mechanical system. The conjugate variable is the charge Q on the capacitor, satisfying the commutation relation [\hat{\phi}, \hat{Q}] = i \hbar. In the phase representation, \hat{Q} = -i \frac{2e}{\hbar} \frac{\partial}{\partial \phi}, where the factor of $2e accounts for charge transport in units of the Cooper pair charge, leading to the number operator \hat{n} = -i \frac{\partial}{\partial \phi} with [\hat{\phi}, \hat{n}] = i. This approach models the circuit as a nonlinear LC oscillator, where the linear capacitance C provides the kinetic energy term and the Josephson junction introduces the nonlinear potential. The classical starting point is the Lagrangian for the undamped RCSJ circuit, derived from the kinetic energy stored in the capacitor and the Josephson potential energy: L = \frac{\hbar^2 C}{8 e^2} \left( \frac{d\phi}{dt} \right)^2 - U(\phi), where U(\phi) = -E_J \cos \phi - E_J \gamma \phi is the tilted washboard potential, with Josephson energy E_J = I_c \hbar / (2e), critical current I_c, bias parameter \gamma = I_b / I_c, and bias current I_b. The factor \hbar^2 C / (8 e^2) arises from the voltage-phase relation V = (\hbar / 2e) d\phi / dt and the capacitive energy (1/2) C V^2. To obtain the Hamiltonian, perform the Legendre transform using the canonical momentum conjugate to \phi, p_\phi = \partial L / \partial (\dot{\phi}) = (\hbar^2 C / 4 e^2) \dot{\phi}, which is proportional to the charge Q = 2 e n. The resulting Hamiltonian is H = \frac{Q^2}{2C} + U(\phi) = 4 E_C \hat{n}^2 - E_J \cos \hat{\phi} - E_J \gamma \hat{\phi}, where the charging energy E_C = e^2 / (2C), and the operators satisfy the required commutation relations upon promotion from classical variables. This form captures the quantum dynamics of the phase qubit, with the bias term tilting the potential to select a metastable well for qubit operation. In the phase qubit regime, the Josephson energy dominates the charging energy, E_J \gg E_C, which localizes the wavefunctions in the potential wells of the washboard, making the phase eigenstates a suitable basis for description. This contrasts with the charge regime (E_C \gg E_J) used in other qubit designs, where charge states are preferred. Under this condition, the system approximates a displaced anharmonic oscillator within each well, with the nonlinearity from the cosine term enabling qubit encoding in the low-lying excited states. The periodic boundary condition \phi \sim \phi + 2\pi reflects the compactness of the phase variable, but for deep wells (large E_J / E_C), tunneling between wells is negligible, allowing an approximation as motion in an infinite potential well for practical calculations.

Energy Levels and Qubit States

The energy levels of the phase qubit are obtained by diagonalizing the Hamiltonian within the confined potential well, yielding a discrete spectrum of metastable states. For low-lying levels, the spectrum can be approximated using a perturbed harmonic oscillator model, accounting for the anharmonicity introduced by the nonlinear . The energy of the m-th level is given approximately by E_m \approx \hbar \omega_p \left( m + \frac{1}{2} \right) - \frac{\chi}{2} m (m - 1), where \omega_p is the small-signal plasma frequency determined by the bias point, and \chi is the anharmonicity parameter arising from higher-order terms in the potential expansion. Alternatively, the full periodic structure of the potential leads to solutions of the Mathieu equation, providing a more complete description of the energy bands and levels. The logical qubit states are encoded in the two lowest energy levels of a selected well: the ground state |0\rangle corresponding to m=0 and the first excited state |1\rangle corresponding to m=1. The qubit transition frequency is f_{01} = (E_1 - E_0)/h \approx 5--$10 GHz, which depends on the bias current I_b through its effect on the well curvature and can be tuned over several hundred MHz. Anharmonicity in the spectrum is quantified by \alpha = (f_{01} - f_{12})/f_{01}, where f_{12} = (E_2 - E_1)/h is the frequency to the second excited state; typical values of \alpha \approx 4\% ensure that microwave drives at f_{01} primarily couple |0\rangle and |1\rangle, minimizing unwanted excitations to higher levels during quantum gate operations. The finite barrier height results in a small tunneling matrix element between the discrete levels and the continuum above the barrier, facilitating coherent quantum dynamics within the well while allowing relaxation via macroscopic quantum tunneling (MQT) or macroscopic action (MA) processes. In early phase qubit implementations, this tunneling limited the energy relaxation time T_1 to approximately 100 ns. Microwave pulses tuned to f_{01} coherently drive transitions between |0\rangle and |1\rangle, producing Rabi oscillations that demonstrate superposition states in the qubit subspace.

Experimental Implementation

Fabrication and Materials

Phase qubits are fabricated using superconducting materials that operate at cryogenic temperatures, typically requiring dilution refrigerators for millikelvin operation. The primary materials include or for the superconducting electrodes, with as the tunnel barrier in ; substrates are commonly or high-resistivity silicon. Dielectrics such as or are used for shunt capacitors, offering low loss tangents (e.g., tan δ ≈ 0.001 for AlOₓ and better for a-Si:H compared to ). Parasitic capacitance from the junction and on-chip elements provides the necessary shunting. Integration with microwave lines for control requires precise alignment to avoid crosstalk, often using a 7-mask process on . The fabrication process involves multi-layer lithographic techniques to create the nonlinear at the core of the phase qubit. Electron-beam lithography defines junction areas of approximately 0.1–1 μm², yielding critical currents (I_c) in the range of 1–10 μA suitable for plasma frequencies around 5–10 GHz. For Al-based junctions, a double-angle shadow evaporation technique deposits the Al-AlOₓ-Al trilayer in situ, with the oxide barrier formed by thermal oxidation (e.g., 10 minutes at 140 mTorr oxygen); Nb-based junctions use DC magnetron sputtering for Nb electrodes and plasma oxidation for the AlOₓ barrier, often in a Nb/Al-AlOₓ/Nb trilayer structure. Additional steps include photolithography for wiring layers, reactive ion etching (e.g., with BCl₃/Cl₂ for Al), and deposition of dielectrics via plasma-enhanced chemical vapor deposition () for a-Si:H. Key challenges in fabrication include achieving junction asymmetry below 1% to maintain coherence, as imbalances introduce unwanted flux and degrade energy levels. Barrier defects from impurities or non-uniform oxidation reduce yield and introduce two-level systems (TLS) that cause decoherence, with TLS density minimized by in-situ oxide growth and anodic isolation layers (~5 nm thick Al or Nb oxide). Critical current uniformity is controlled via oxidation time and Ar ion milling to clean interfaces, but scaling remains limited by these material imperfections and the need for cryogenic shielding. Over time, fabrication has evolved from manual processes in early demonstrations around 2002–2003, which relied on basic shadow evaporation for proof-of-concept devices, to more automated, semi-CMOS-compatible methods by the 2010s using commercial foundries for trilayer junctions. These advances improved coherence times from ~8 ns to over 600 ns through smaller junctions and low-loss dielectrics, but scaling to more than 10 qubits has been constrained by yield issues and integration complexity. Following the 2010s, experimental focus shifted to and other designs, with minimal further development on phase qubits due to challenges in scaling and coherence.

Biasing, Control, and Readout

Phase qubits are biased using a direct current (DC) supplied through on-chip transmission lines or, in the case of dc SQUID implementations, via the SQUID loop itself to provide inductive isolation from low-frequency noise. The bias current I_b is set just below the critical current I_0, typically tuning the potential barrier height \Delta U to approximately 1-5 kT (where k is and T is the operating temperature around 20 mK), which optimizes the energy relaxation time T_1 by balancing thermal escape rates and quantum tunneling. Control of the qubit state is achieved primarily through microwave pulses delivered via on-chip striplines, with frequencies in the 5-10 GHz range corresponding to the transition energy between the ground and first excited states. These pulses, typically lasting 10-50 ns, drive \pi/2 or \pi rotations for or , with the Rabi frequency determined by the pulse amplitude. In loop-based configurations, such as dc phase qubits, additional fine-tuning of the qubit frequency is performed by applying a flux bias through the superconducting loop, enabling adjustable coupling to other qubits or resonators. To mitigate dephasing from low-frequency noise, sequences are employed, involving a \pi pulse midway between two \pi/2 pulses to refocus phase errors, extending the coherence time T_2. Readout relies on a switching measurement technique, where the bias current is rapidly increased to the escape point of the tilted washboard potential, causing the junction to switch to a finite-voltage state if the qubit is in the due to enhanced tunneling probability. The resulting voltage snap, on the order of 1 mV across the junction, is detected in a single-shot manner using a sensitive amplifier, such as a dc SQUID or a high-electron-mobility transistor (HEMT) at 4 K, followed by room-temperature amplification. Early demonstrations in the 2000s achieved single-shot readout fidelities around 90%, with the ground state showing near-unity detection (>99%) and the around 85%. For two-qubit operations, phase qubits are coupled capacitively via an inter-qubit or through shared resonators, enabling like the iSWAP, which swaps excitations between qubits with an imaginary phase factor. A seminal experiment demonstrated coherent oscillations between coupled phase qubits with a coupling strength of about 80 MHz, achieving a two-qubit gate fidelity of approximately 70%, limited primarily by measurement crosstalk and relaxation. These interactions form the basis for quantum when combined with single-qubit rotations.

Performance Characteristics

Advantages

Phase qubits offer significant tunability through a simple current bias applied to the Josephson junction, which adjusts the qubit's transition frequency over a range of approximately 7-10 GHz without introducing flux noise issues associated with tuning in other superconducting qubit designs. This bias current directly modulates the effective potential barrier height in the tilted washboard potential, enabling precise control of the plasma frequency and thus the qubit energy levels. The strong anharmonicity inherent to phase qubits, arising from the nonlinear Josephson inductance, facilitates fast quantum gate operations with typical single-qubit gate times on the order of 10-20 ns via direct microwave driving at the transition frequency. This rapid execution stems from the large Josephson energy E_J relative to the charging energy E_C, allowing selective addressing of the |0⟩ to |1⟩ transition without significant leakage to higher levels. A key strength of phase qubits is their insensitivity to charge noise, achieved through a large E_J / E_C ratio (typically exceeding 1000), which renders the phase variable—the conjugate to charge—robust against fluctuations in the number of Cooper pairs, in contrast to charge-sensitive qubits. This design choice minimizes from offset charges at the junction electrodes, enhancing the stability of qubit operations in practical cryogenic environments. Readout in phase qubits is notably fast, relying on deterministic macroscopic quantum tunneling where a sudden increase in bias current causes the to escape the within approximately 100 ns, outperforming the integration times required for dispersive readout in qubits. This switching-based method provides a high-contrast voltage signal for state discrimination, enabling efficient measurements. Early demonstrations of phase qubits achieved some of the highest single-qubit fidelities reported at the time, reaching approximately 99% by 2008, underscoring their potential for high-performance quantum control.

Limitations and Decoherence

Phase qubits suffer from several decoherence mechanisms that limit their operational fidelity and coherence times. The primary relaxation process for the is macroscopic quantum tunneling (MQT) or thermal activation (MA) escape from the tilted washboard potential, with relaxation times T_1 typically ranging from 100 to 500 depending on junction size and materials. , characterized by transverse relaxation times T_2, is dominated by low-frequency 1/f noise in flux and current, yielding T_2 values of 50-200 , though spin-echo techniques can extend this to around 300 . The qubit's high sensitivity to environmental noise exacerbates these limits. Fluctuations in the bias current (\delta I \approx 1 nA) can induce significant dephasing by shifting the energy splitting, as the qubit frequency varies strongly with bias near the optimal point. Additional loss channels include thermal photons exciting the qubit and quasiparticle tunneling across the Josephson junction, both contributing to reduced T_1. Dielectric losses from two-level systems (TLS) in shunt materials, such as amorphous oxides, further degrade coherence, with loss tangents around $10^{-3} to $10^{-2}. Scalability poses significant challenges for phase qubit arrays. Fixed capacitive coupling between qubits leads to unwanted crosstalk, with interaction probabilities up to 20%, complicating multi-qubit control. Readout, often via MQT detection, disturbs the qubit state, resulting in measurement errors around 10%. By 2010, coherence times in phase qubits had plateaued at approximately 1 μs, constrained by these persistent noise sources. Efforts to mitigate decoherence include dynamical via echo pulses to refocus from 1/f and improved low-pass filtering to suppress bias current fluctuations. Material optimizations, such as using low-loss dielectrics like hydrogenated , have also extended T_1 beyond 500 ns in select devices.

Comparison to Other Qubits

Phase qubits differ from charge qubits primarily in their reduced sensitivity to charge , owing to the large junction that delocalizes the phase wavefunction across many charge states, whereas charge qubits, operating in the regime where charging energy E_C dominates Josephson energy E_J (E_J / E_C \ll 1), exhibit high susceptibility to offset charge fluctuations, leading to rapid . Both designs suffer from comparably short coherence times, with phase qubits achieving T_2 up to approximately 300 ns via spin-echo techniques, similar to the few-nanosecond times observed in early charge qubits without . Charge qubits were rendered obsolete earlier due to their noise challenges, while phase qubits provided a transitional design before more robust architectures emerged. In comparison to flux qubits, phase qubits share a phase-based operating regime with high E_J / E_C ratios but employ a single current-biased Josephson junction rather than the superconducting quantum interference device (SQUID) loop structure that enables s in flux qubits. This results in simpler switching-based readout for phase qubits, achieving high fidelity (up to 96%) in tens of nanoseconds, whereas flux qubits rely on more complex inductive tuning and states for operation. However, phase qubits demonstrate poorer rejection of flux noise compared to flux qubits tuned to their symmetry points, where first-order insensitivity can extend coherence, though both remain vulnerable to low-frequency 1/f noise sources. Transmon qubits represent a significant from qubits, featuring a shunted Josephson junction with E_J / E_C \gg 1 (typically >50) that exponentially suppresses charge noise sensitivity while preserving sufficient for operation, unlike the current-biased qubit which requires operation near the critical current to achieve similar delocalization but introduces greater exposure to and critical-current fluctuations. Transmons exhibit longer relaxation times (T_1 \sim 100 \, \mus in optimized designs) and improved scalability through planar fabrication without dc bias lines, enabling denser arrays, whereas qubits, despite faster macroscopic quantum tunneling rates for rapid , are more noise-prone and require complex biasing that hinders integration. Modern designs like fluxonium and cat qubits build on phase qubit principles by incorporating large inductors or driven nonlinear oscillators to further mitigate , providing coherent protection against and errors that phase qubits lack due to their direct exposure to junction parameter fluctuations. Fluxonium qubits, for instance, achieve millisecond-scale T_1 times through inductive shunting that suppresses sensitivity, while cat qubits use continuous-variable encoding to correct bit- and -flip errors more effectively than the discrete-state phase qubit. Although phase qubits informed these advancements in macroscopic quantum tunneling and strong , they do not incorporate the error-suppressing mechanisms central to these newer architectures. As of 2025, phase qubits are rarely employed in scalable , with leading efforts by and favoring transmon-based processors for their superior and , though phase qubits retain niche utility in studies of macroscopic quantum tunneling phenomena.

References

  1. [1]
    [PDF] Superconducting Phase Qubits - UCSB Physics
    Superconducting phase qubits are a type of qubit with a large capacitance, allowing strong coupling and a 50Ω impedance, and are used for quantum computation.
  2. [2]
  3. [3]
    [PDF] Superconducting Qubits and the Physics of Josephson Junctions
    Implementing the phase qubit is challenging because a current bias is required with large impedance. This impedance requirement can be met by biasing the ...
  4. [4]
    Simulating the fabrication of aluminium oxide tunnel junctions - Nature
    Jan 28, 2021 · Superconducting quantum computers often use aluminium oxide tunnel junctions as Josephson junctions to introduce the required nonlinearity1,2,3, ...
  5. [5]
    [PDF] Chapter 1 JOSEPHSON JUNCTION MATERIALS RESEARCH ...
    This procedure leads to the preferred oxide thickness, giving junction critical-current densities of ∼ 15 − 20 A/cm2, appropriate for our phase qubits. The base ...
  6. [6]
    Algorithms for automatic measurement of SIS-type hysteretic ...
    Some electrical parameters of the SIS-type hysteretic underdamped Josephson junction (JJ) can be measured by its current-voltage characteristics (IVCs).
  7. [7]
    Observation of High Coherence in Josephson Junction Qubits ...
    Dec 5, 2011 · The observed phase coherence factor, Q 2 ∼ 700 000 , is an order of magnitude larger than previous superconducting devices [10, 25] . Since ...
  8. [8]
    CURRENT‐VOLTAGE CHARACTERISTICS OF JOSEPHSON ...
    A model for superconducting Josephson junctions analogous to the driven simple pendulum with damping gives dc I‐V curves displaying hysteresis for light ...
  9. [9]
  10. [10]
    [PDF] superconducting josephson - phase quantum bits - Uni Ulm
    phase qubits, and likewise other superconducting qubit approaches, are promising can- didates to realize a solid-state quantum computer. However, their ...<|control11|><|separator|>
  11. [11]
    [PDF] Quantum Bits with Josephson Junctions - Franco Nori
    Roughly speaking, the charge qubit is a box for charge, controlled by an external voltage Vg; the flux qubit is a loop controlled by an external magnetic flux ...
  12. [12]
    Effects of anharmonicity of current-phase relation in Josephson ...
    Apr 1, 2015 · It was demonstrated that splitting of energy levels in phase qubit as well as energy gap in charge qubit reveals similar behavior with energy ...
  13. [13]
    Superconducting phase qubits | Quantum Information Processing
    Feb 18, 2009 · Abstract. Experimental progress is reviewed for superconducting phase qubit research at the University of California, Santa Barbara. The phase ...Missing: paper | Show results with:paper
  14. [14]
    Superconducting Qubit Storage and Entanglement with ...
    Aug 10, 2004 · ... coherence time of order 100 ns, a time already demonstrated in the phase qubit [18] . More extensive operations could be performed with a ...
  15. [15]
    Rabi Oscillations in a Large Josephson-Junction Qubit
    Aug 21, 2002 · We have designed and operated a circuit based on a large-area current-biased Josephson junction whose two lowest energy quantum levels are used to implement a ...
  16. [16]
    None
    ### Summary of Fabrication Process and Materials for Phase Qubits with Trilayer Junctions
  17. [17]
    [PDF] Fabrication and measurements of hybrid Nb/Al Josephson junctions ...
    May 2, 2014 · SFS phase shifters have already been successfully implemented in Nb-based phase qubit circuits [18], but they haven't yet been used in flux ...
  18. [18]
    Tripartite interactions between two phase qubits and a resonant cavity
    Aug 1, 2010 · Observation of quantum oscillations between a Josephson phase qubit and a microscopic resonator using fast readout. Phys. Rev. Lett. 93 ...Missing: Duffing | Show results with:Duffing
  19. [19]
    [PDF] arXiv:0805.0164v1 [cond-mat.supr-con] 1 May 2008
    May 1, 2008 · These are phase, flux and charge qubits. The relation between the parameters hω0,EC and EJ and the way these qubits are biased distinguishes ...
  20. [20]
    Decoherence in Josephson Qubits from Dielectric Loss
    Dielectric loss from two-level states is shown to be a dominant decoherence source in superconducting quantum bits.Abstract · Article Text
  21. [21]
    Quantum Bits with Josephson Junctions - SpringerLink
    Sep 17, 2019 · When E_J / E_C \gg 1, the opposite holds. This is the case for both the flux qubit and the phase qubit. 17.3.1 Charge Qubit. We now discuss the ...
  22. [22]
    Dynamical Autler-Townes control of a phase qubit | Scientific Reports
    Sep 10, 2012 · The results demonstrate that the qubit can be used as a ON/OFF switch with 100 ns operating time-scale for the reflection/transmission of ...
  23. [23]
    High-Fidelity Gates in a Single Josephson Qubit | Phys. Rev. Lett.
    Here, we measure the fidelity of a single qubit gate for a Josephson phase qubit, demonstrating substantial progress towards this goal. Using the new ...
  24. [24]
    [PDF] Decoherence of a superconducting qubit due to bias noise
    Mar 25, 2003 · The main purpose of this paper is to give a physical picture of decoherence, showing that it can be understood as a random fluctuation in the ...
  25. [25]
    [PDF] Materials Origins of Decoherence in Superconducting Qubits
    Sep 30, 2009 · Abstract—Superconducting integrated circuits incorporating. Josephson junctions are an attractive candidate for scalable.<|control11|><|separator|>
  26. [26]
    A quantum engineer's guide to superconducting qubits
    Jun 17, 2019 · One remarkable feature of superconducting qubits is that their energy-level spectra are governed by circuit element parameters and thus are ...
  27. [27]
  28. [28]
    Superconducting quantum computers: who is leading the future?
    Aug 19, 2025 · Single-qubit gates achieved fidelity up to 99.9% with gate durations as short as 13.3 ns , demonstrating stability over extended periods [60].