Fact-checked by Grok 2 weeks ago

Transmon

The transmon is a superconducting artificial used as a quantum bit () in and (QED), consisting of a Josephson junction shunted by a large to form a weakly anharmonic nonlinear . It operates by quantizing the charge and phase across the junction, with energy levels that can be controlled via microwave pulses for qubit manipulation. Developed in 2007 as an improvement over the charge-sensitive box (CPB) qubit, the transmon achieves charge insensitivity by operating at a high of Josephson E_J (typically 20–30 GHz) to charging E_C (around 0.35 GHz), exponentially suppressing sensitivity to charge noise while preserving sufficient (100–300 MHz) for selective addressing. This design eliminates the need for precise charge bias tuning at "sweet spots," enhancing scalability for multi- systems compared to earlier superconducting qubits, which suffered from times limited to nanoseconds due to environmental fluctuations. In practice, transmons are fabricated on or substrates using planar superconducting circuits, often incorporating a loop for in-situ frequency tuning via external , with operating frequencies tunable in the 4–10 GHz range. Readout is typically performed dispersively through a coupled resonator, enabling high-fidelity state detection. By 2021, advanced transmon implementations achieved coherence times exceeding 300 μs. As of 2025, further advancements have extended coherence times beyond 1 ms, supporting fault-tolerant and applications in quantum simulation, sensing, and information processing.

Background

Superconducting Qubits

Superconducting qubits are artificial atoms implemented in solid-state circuits made from superconducting materials and Josephson junctions, which allow for the encoding of in discrete energy states. These devices exploit the macroscopic of superconductors at millikelvin temperatures to mimic the behavior of natural atoms, but with the advantages of lithographic fabrication and tunability. The Josephson junction, consisting of two superconductors separated by a thin insulating barrier, provides the essential nonlinearity required to create an anharmonic energy spectrum for qubit operations. The basic operation of superconducting qubits relies on the quantization of electromagnetic oscillations in LC circuits, where the circuit's degrees of freedom—charge and flux—become operators in the quantum regime, leading to well-defined energy levels that encode the qubit states. In the simplest case of a linear LC oscillator, the Hamiltonian is H = \frac{Q^2}{2C} + \frac{\Phi^2}{2L}, where Q is the charge operator, \Phi is the flux operator, C is the , and L is the ; this model yields equally spaced energy levels, but the addition of a Josephson junction introduces nonlinearity to select computational states. Microwave pulses are used to drive transitions between these levels, enabling single- and multi-qubit gates essential for quantum computation. Superconducting qubits gained prominence in the early 2000s as scalable for , following initial demonstrations of in the late 1990s, due to their integration with planar microwave technology and potential for on-chip control. This approach offers advantages over other qubit modalities, including fast gate times on the order of nanoseconds and compatibility with fabrication processes. The transmon is a specific type of charge-based superconducting designed for enhanced .

Cooper Pair Box

The Cooper pair box (CPB) is a foundational design for superconducting charge qubits, consisting of a small superconducting island connected to a larger superconducting reservoir via a single Josephson junction, with an additional gate capacitor coupled to the island for external control. The qubit states are encoded in the number of excess Cooper pairs residing on the isolated island, where each Cooper pair carries a charge of -2e. In the superconducting state, the island hosts a macroscopic quantum state characterized by the integer number operator \hat{n} of these pairs, and the Josephson junction enables coherent tunneling of Cooper pairs between the island and the reservoir. This setup was first experimentally realized and demonstrated to exhibit quantum coherence in 1999. The CPB operates in the charge regime, where the charging energy E_C significantly exceeds the Josephson coupling energy E_J (typically E_C \gg E_J), making the excess charge on the island the dominant degree of freedom. A gate voltage V_g applied through the gate capacitor with capacitance C_g induces an offset charge on the island, parameterized by the dimensionless gate charge n_g = C_g V_g / (2e). By tuning n_g to a degeneracy point (e.g., n_g \approx 0.5), the energy levels of adjacent charge states, such as |n=0\rangle and |n=1\rangle, become nearly degenerate, allowing the formation of a |\psi\rangle = \alpha |0\rangle + \beta |1\rangle through coherent tunneling mediated by the Josephson junction. This two-level system can then be manipulated via pulses or voltage gates to perform quantum operations. A primary limitation of the CPB arises from its high sensitivity to charge noise, stemming from fluctuating background charges (e.g., trapped defects or impurities) that randomly shift n_g, causing rapid dephasing of the superposition states. This charge dispersion leads to short coherence times, typically less than 1 μs in early experiments, which severely restricts the fidelity of quantum operations and scalability. The energy levels of the CPB are described by quantizing the classical circuit Hamiltonian. The classical charging energy arises from the electrostatic energy stored on the island's total capacitance C_\Sigma = C_J + C_g, where C_J is the Josephson junction capacitance, giving the charging term E_{ch} = \frac{(2e (n_g - n))^2}{2 C_\Sigma} = 4 E_C (n_g - n)^2, with E_C = e^2 / (2 C_\Sigma) as the charging energy and n the number of excess Cooper pairs. The Josephson junction contributes a nonlinear inductive potential E_J (1 - \cos \phi), where \phi is the superconducting phase difference across the junction and E_J = I_c \Phi_0 / (2\pi) is the Josephson energy (I_c critical current, \Phi_0 = h/(2e) flux quantum). Thus, the classical Hamiltonian is H_{cl} = 4 E_C (n_g - n)^2 + E_J (1 - \cos \phi). To quantize, treat the phase \phi and charge number n as canonically conjugate variables satisfying [\hat{\phi}, \hat{n}] = i, analogous to position and momentum. In the charge basis |n\rangle, the charging term is diagonal: \langle n' | 4 E_C (\hat{n} - n_g)^2 | n'' \rangle = 4 E_C (n - n_g)^2 \delta_{n',n''}, while the Josephson term introduces off-diagonal coupling: \langle n' | -E_J \cos \hat{\phi} | n'' \rangle = -(E_J/2) (\delta_{n',n''+1} + \delta_{n',n''-1}), since \cos \hat{\phi} = (\hat{\phi}^i + \hat{\phi}^{-i})/2 and \exp(i \hat{\phi}) |n\rangle = |n+1\rangle. The full quantum Hamiltonian is therefore \hat{H} = 4 E_C (\hat{n} - n_g)^2 - E_J \cos \hat{\phi}, with energy eigenvalues obtained by diagonalizing in the truncated charge basis (typically two levels near degeneracy). The constant shift E_J in the potential is often omitted as it does not affect dynamics. This model captures the anharmonicity essential for qubit control.

History

Invention

The transmon qubit was proposed in 2007 by Jens Koch, A. Houck, and colleagues at as a solution to the pronounced sensitivity to charge noise that plagued earlier superconducting charge qubits like the Cooper pair box (CPB). This motivation stemmed from the CPB's operation in a regime where the Josephson energy E_J and charging energy E_C were comparable (E_J \approx E_C), leading to rapid dephasing times on the order of nanoseconds due to fluctuating offset charges. The core innovation of the transmon involved modifying the CPB circuit by adding a large shunt in parallel with the Josephson junction, which significantly increases the total and thereby reduces E_C (since E_C = e^2 / 2C), allowing operation in the regime where E_J \gg E_C (typically ratios of 50 or higher). This adjustment exponentially suppresses the charge dispersion of the energy levels— the variation in transition frequency with offset charge—while the decreases only as a weak power law, preserving sufficient qubit-photon coupling for applications. The design was detailed in the seminal paper "Charge-insensitive qubit design derived from the Cooper pair box," published in Physical Review A (volume 76, issue 4, article 042319). Theoretical predictions in the 2007 paper indicated that this reduced charge sensitivity could extend times (T_2) to hundreds of microseconds, a substantial improvement over the CPB's typical ~1 μs limit, by making the nearly immune to 1/f charge noise. Early experimental validation came in 2008, when the same Yale group fabricated and measured the first transmon devices, achieving relaxation times (T_1) of approximately 1.9 μs and times (T_2^*) up to 2.2 μs at the charge-insensitive sweet spot, confirming and a marked enhancement over prior CPB limited to a few nanoseconds. These results demonstrated the transmon's potential for scalable architectures.

Key Milestones

Following the invention of the transmon qubit by researchers at in 2007, subsequent experimental efforts rapidly advanced its implementation in multi-qubit systems. Between 2008 and 2010, pioneering demonstrations of multi-qubit operations emerged, notably from the Yale team, which in 2009 realized a two-qubit superconducting using transmon qubits coupled via a bus, achieving entangling gates and quantum algorithms with fidelities exceeding 80%. Around the same period, the UC Santa Barbara group contributed to early scalable architectures by exploring in superconducting qubits, laying groundwork for transmon-based two-qubit gates. During the 2010s, major industry players adopted the transmon as the cornerstone of their platforms; IBM integrated transmons into its early quantum processors starting around 2012, while began deploying them in scalable arrays by 2015, emphasizing fixed-frequency designs for improved . A landmark achievement came in 2019 with 's , featuring 53 transmon qubits, which demonstrated by completing a random circuit sampling task in 200 seconds—a estimated to take classical supercomputers 10,000 years. Key performance milestones included a record coherence time of approximately 100 μs for a transmon qubit coupled to a three-dimensional () superconducting cavity, enabling more robust quantum operations. Integration of transmons with cavities, first demonstrated around 2011, provided enhanced electromagnetic isolation from , boosting coherence times and facilitating multi-qubit entanglement with reduced . From 2023 to , progress accelerated toward scalability; outlined a roadmap targeting modular systems exceeding 4,000 transmon qubits by to enable error-corrected quantum utility. In , a team led by Andrew Houck introduced a tantalum-on- transmon that achieved millisecond-scale coherence times—over 1,000 times the reliability of early transmons—through minimized in the high-resistivity silicon substrate.

Theoretical Model

Circuit Quantization

The quantization of the transmon circuit involves treating the classical electrical network—a Josephson junction in parallel with a shunt capacitor—as a quantum mechanical system. The key variables are the gauge-invariant phase difference \phi across the junction and the conjugate charge Q on the capacitor plates. These are promoted to operators \hat{\phi} and \hat{Q} satisfying the commutation relation [\hat{\phi}, \hat{Q}] = i (2e), where $2e is the elementary charge of a Cooper pair, reflecting the discrete nature of charge transport in superconductors. The Josephson junction introduces a nonlinear inductance through its characteristic relations: the supercurrent I = I_c \sin \phi and the voltage-phase relation V = (\hbar / 2e) \dot{\phi}. This phase-dependent inductance L_J(\phi) = (\hbar / 2e I_c) / \cos \phi provides the anharmonicity necessary for selective addressing of states, distinguishing the transmon from oscillators. The quantum description begins with the classical for the , formulated in terms of the \phi as the dynamical variable. For the shunted junction, the is L = \frac{C}{2} \dot{\phi}^2 + E_J \cos \phi, where C is the total shunt capacitance, \dot{\phi} = d\phi / dt, and E_J = I_c \hbar / 2e is the Josephson energy scale. The first term captures the capacitive , while the second derives from the Josephson V(\phi) = -E_J \cos \phi, which combines with the quadratic charging contribution to form the landscape. The conjugate momentum is the charge Q = \partial L / \partial \dot{\phi} = C \dot{\phi}. The classical Hamiltonian follows from the Legendre transform H = Q \dot{\phi} - L, yielding H = \frac{Q^2}{2C} - E_J \cos \phi. Quantizing this expression gives the operator Hamiltonian \hat{H} = \frac{\hat{Q}^2}{2C} - E_J \cos \hat{\phi}. Equivalently, defining the reduced charge operator \hat{n} = \hat{Q} / (2e) with [\hat{\phi}, \hat{n}] = i, and the charging energy E_C = e^2 / (2C), the Hamiltonian becomes \hat{H} = 4 E_C \hat{n}^2 - E_J \cos \hat{\phi}. This framework maps the circuit to a quantum anharmonic oscillator, where the nonlinear cosine term perturbs the otherwise spectrum.

Hamiltonian

The effective of the transmon , derived from the quantization of its superconducting circuit, is given by H = 4 E_C (n - n_g)^2 - E_J \cos \phi, where E_C is the charging , E_J is the Josephson , \phi is the phase difference across the Josephson junction, n is the number operator conjugate to \phi satisfying [ \phi, n ] = i, and n_g is the dimensionless gate-induced charge offset. In the regime of large Josephson coupling where E_J / E_C \gg 1, perturbative analysis approximates the transmon as a weakly anharmonic oscillator. The lowest two levels form an effective two-level system with transition frequency \omega_{01} \approx \sqrt{8 E_C E_J} - E_C and anharmonicity \alpha \approx -E_C, where the negative anharmonicity enables selective addressing of transitions. The charge dispersion, which quantifies the sensitivity of energy levels to fluctuations in n_g, is exponentially suppressed in this regime: \varepsilon(n_g) \approx e^{-\sqrt{8 E_J / E_C}} \sqrt{2 E_C^3 / E_J}. This term introduces a small periodic to the energy spectrum, but its amplitude decreases rapidly with increasing E_J / E_C, rendering the transmon robust against charge noise. For higher energy levels, the associated with the is [4 E_C (-i \frac{d}{d \phi} - n_g)^2 - E_J \cos \phi ] \psi(\phi) = E \psi(\phi). For n_g = 0, it reduces to a Mathieu -4 E_C \frac{d^2 \psi}{d \phi^2} - E_J \cos \phi \, \psi = E \psi. The solutions confirm the transmon's behavior as a weakly anharmonic oscillator, with level spacings approaching those of a while retaining sufficient for multi-level operations.

Design Principles

Shunt Capacitance

The shunt capacitance in the transmon qubit is a large capacitor connected in parallel with the Josephson junction, forming the primary modification from the Cooper pair box design. This configuration increases the total capacitance C_\Sigma = C_J + C_{sh}, where C_J is the junction capacitance and C_{sh} is the shunt capacitance, thereby reducing the charging energy E_C = e^2 / (2 C_\Sigma) while the Josephson energy E_J remains fixed by the junction critical current. The primary purpose is to suppress sensitivity to charge noise by operating in the regime where E_J \gg E_C, exponentially diminishing the charge dispersion of energy levels. Typically, C_{sh} \gg C_J, with total capacitances C_\Sigma on the order of 70–100 , yielding E_J / E_C ratios of 50–100 that optimize charge insensitivity without excessive loss of . These ratios ensure the transmon's energy levels are nearly insensitive to offset charge fluctuations, a key improvement over charge . Geometrically, the shunt capacitance is realized on-chip using planar structures such as interdigital capacitors or segments, which provide the necessary fF-scale while minimizing the physical size of the Josephson to reduce extrinsic losses. These designs, often featuring finger-like or cross-shaped electrodes, allow for precise control of without increasing junction area, facilitating in scalable circuits. A key trade-off arises with larger C_{sh}: while it enhances noise immunity through reduced charge , it slightly diminishes the of the transmon's energy spectrum, which scales approximately as (E_J / E_C)^{-1/2}, potentially complicating selective addressing of higher levels for multi-photon operations. Nonetheless, ratios above 50 maintain sufficient for practical control.

Josephson Junction Characteristics

The Josephson junction is the primary source of nonlinearity in the transmon qubit, contributing the inductive Josephson energy E_J that creates an anharmonic potential essential for distinguishing states. This nonlinearity arises from the junction's characteristic cosine potential, which ensures that the energy spacing between the and first excited states differs from that between the first and second excited states, enabling precise control of the 0–1 transition without inadvertently populating higher levels. The Josephson energy is defined as E_J = \frac{I_c \Phi_0}{2\pi}, where I_c denotes the critical current of the junction and \Phi_0 = h/(2e) is the magnetic flux quantum. In standard transmon implementations, aluminum-based Al/AlO_x/Al tunnel junctions are employed, featuring areas approximately 0.01 \mum^2 to yield E_J/h \sim 10–30 GHz, with the oxide tunneling barrier typically around 1 nm thick to control the critical current density. These specifications balance the desired anharmonicity while minimizing charge sensitivity in the transmon regime. For enhanced flexibility, variations incorporate a superconducting quantum interference device (SQUID) in place of a single junction, where two parallel Josephson junctions form a loop that allows E_J to be tuned via an applied threading the loop. This flux-tunable configuration, with E_J modulated as E_J = E_{J,\max} |\cos(\pi \Phi / \Phi_0)|, facilitates adjustable frequencies or coupling strengths in multi-qubit systems. The junction integrates with a shunt to set the overall dynamics, prioritizing the nonlinear inductive response over linear capacitive effects.

Fabrication

Materials and Structure

Transmon qubits are constructed using superconducting metals such as or for the electrodes and interconnecting wiring, which provide low-resistance paths for supercurrents at millikelvin temperatures. The Josephson junctions at the core of the device typically employ an Al/AlO_x/Al trilayer structure, where the thin AlO_x tunnel barrier is formed by controlled oxidation of an layer, enabling weak-link with critical currents on the order of microamperes. These materials are chosen for their compatibility with standard techniques and ability to maintain below their critical temperatures (1.2 K for and 9.2 K for ). The overall structure adopts a planar on a high-purity , which offers low losses compared to alternatives like . Superconducting ground planes flank the central in a layout, minimizing unwanted modes and radiation losses. The transmon consists of two large superconducting islands forming the shunt , interconnected by the Josephson ; one island is capacitively coupled to a nearby readout for dispersive measurement. This design ensures the qubit's nonlinear is shunted by a geometric , suppressing charge sensitivity. Typical dimensions include a Josephson junction width of approximately 100 nm to achieve the desired tunneling resistance, while the shunt capacitor pads extend to about 200 μm across, yielding a shunt C_sh of 50–100 fF that dominates the total charging energy. The junction material properties directly influence the Josephson energy E_J, tuning the qubit's . Recent advances in 2025 have introduced tantalum-based films on high-resistivity (>20 kΩ·cm) substrates for transmon fabrication, dramatically reducing bulk substrate losses relative to earlier tantalum-on-sapphire devices and enabling times exceeding 1 ms—over three times longer than prior benchmarks.

Manufacturing Process

The manufacturing process for transmon qubits begins with the preparation of a superconducting , typically or , followed by the deposition of materials such as for low-loss transmission lines. is employed to define the nanoscale features of the Josephson junctions, using a bilayer resist like PMMA on to create suspended bridges that enable shadow . This step achieves sub-100 resolution critical for junction dimensions around 100-200 nm². The core of the junction formation involves double-angle evaporation in a high-vacuum chamber to deposit the aluminum electrodes. First, aluminum is evaporated at normal incidence to form the bottom electrode, followed by in-situ oxidation to create the AlOₓ tunnel barrier through exposure to diluted oxygen at controlled pressure (typically 0.001-0.1 mbar) and temperature. A second evaporation at an angled tilt (around 30-60°) deposits the top aluminum electrode, overlapping the bottom layer to complete the Al/AlOₓ/Al structure while defining the junction area via the shadow effect of the resist bridge. This technique ensures precise control over critical current densities, adjustable via oxidation parameters to tune junction inductance. Post-fabrication processing includes lift-off in solvents to remove excess metal and resist, followed by additional patterning steps for capacitors and interconnects using optical or electron-beam lithography and reactive ion etching. Wafers are then diced into chips using a protective resist coating to prevent damage, with subsequent cleaning via acetone, isopropanol, and oxygen plasma ashing. Chips are wire-bonded or bump-bonded to interposer substrates for electrical connections, then packaged in shielded enclosures and mounted within dilution refrigerators operating below 20 mK to suppress thermal noise. Yield challenges in transmon fabrication primarily stem from defects, such as pinholes or non-uniform oxidation, requiring defect rates below 1% for viable multi-qubit devices; this is achieved through optimized angles and thicknesses to minimize below 1 nm RMS. Shadow masks, often free-standing structures, are used to avoid resist contamination and organic residues during metal deposition, improving reproducibility and reducing critical current variations to under 4%. For scalability, CMOS-compatible processes have enabled the production of chips with over 100 qubits, as demonstrated by in 2024 with 300 mm wafer-scale fabrication yielding over 98% functional qubits across 400-device arrays through automated optical and ; adopted 300 mm in 2025, paving the way for thousand-qubit systems.

Operation

Control and Readout

Control of transmon qubits is achieved by applying resonant pulses to drive transitions between the qubit's energy levels, typically at frequencies around 5 GHz corresponding to the |0⟩ to |1⟩ transition. These pulses induce Rabi oscillations, enabling single-qubit operations such as π-pulses that coherently flip the qubit state from ground to excited or vice versa. In flux-tunable transmon designs, where the Josephson junction is replaced by a superconducting quantum interference device (), additional control is provided via DC flux lines that thread through the SQUID loop, allowing dynamic adjustment of the qubit frequency over a range of several hundred MHz to facilitate multi-qubit interactions or error mitigation. Readout of the transmon state relies on dispersive coupling to a high-quality-factor , where the 's presence shifts the 's in a state-dependent manner by a dispersive shift χ of approximately 10 MHz—the is higher when the is in |0⟩ and lower when in |1⟩. This shift is detected through homodyne of the signal transmitted through or reflected from the , with a probe tone applied near the (typically 6-7 GHz). The readout process is quantum nondemolition, preserving the state during , and achieves assignment fidelities exceeding 99% with integration times as short as 100 ns using optimized pulse shapes and cryogenic amplification. Characterization of control fidelity involves sequences such as Rabi experiments, where the duration of a pulse is varied to observe coherent oscillations, and Ramsey experiments, which apply two π/2-pulses separated by a free-evolution period to assess phase stability. These sequences confirm high-fidelity operations with minimal errors from pulse imperfections. To mitigate the , where the relaxes prematurely through the low-impedance readout channel, high-impedance bandpass or low-pass Purcell filters are integrated into the readout line, suppressing emission at the qubit frequency while passing the higher-frequency resonator signal and extending coherence during measurement.

Multi-Qubit Coupling

Multi-qubit in transmon systems enables the implementation of entangling quantum essential for quantum circuits, primarily through , inductive, or resonator-mediated interactions between qubits. connects adjacent transmons via a shared between their voltage antinodes, facilitating transverse interactions that realize iSWAP , where excitations are swapped with an added π shift. The strength g/2\pi typically ranges from 20 to 100 MHz, depending on the coupling and qubit geometry, allowing gate times on the order of tens of nanoseconds. This fixed is well-suited for nearest-neighbor connectivity in planar architectures but introduces always-on interactions that require careful pulse to mitigate unwanted ZZ terms. Inductive coupling via a superconducting quantum interference device () provides tunable interactions by flux-biasing the loop in a coupler element, modulating the effective mutual between transmons and enabling control over ZZ interactions for controlled-phase (CZ) gates. The ZZ coupling strength can be tuned from near zero to several MHz by adjusting the through the , suppressing residual interactions when off and activating entangling operations on demand. This approach reduces in multi-qubit arrays by allowing selective activation, though it demands precise control to avoid decoherence from flux noise. Bus resonators mediate fixed-frequency coupling between transmons through capacitive connections to a common microwave , enabling effective all-to-all interactions via virtual exchange in the dispersive regime. Typical - coupling strengths g/2\pi are 100–200 MHz, supporting iSWAP-like without direct qubit-qubit overlap and facilitating scalable processor designs. The acts as a communication bus, with detunings ensuring low population during operations. Two-qubit gate fidelities in transmon systems routinely exceed 99%, achieved through optimized control pulses incorporating dynamical decoupling sequences like XY4 to suppress and crosstalk. Challenges persist in crosstalk suppression, particularly from spectator qubits inducing unwanted interactions, which dynamical decoupling mitigates by averaging out error channels during execution.

Performance

Coherence and Decoherence

The of transmon qubits is characterized by two primary timescales: the energy relaxation time T_1, which measures the from the to the , and the time T_2, which quantifies the loss of phase information. Typical T_1 values for transmon qubits now range from 100 to 300 μs (as of 2025), primarily limited by energy relaxation through losses in the . Meanwhile, T_2 values are typically 50 to 150 μs (as of 2025), with often limited by residual low-frequency noise sources including 1/f charge and noise. The main mechanism for T_1 decoherence is in the shunt capacitor, where energy dissipates through non-radiative processes quantified by the loss tangent \tan \delta \approx 10^{-6}, often originating from interfaces and amorphous dielectrics. Additionally, two-level (TLS) defects at metal-dielectric interfaces act as resonant absorbers, contributing significantly to both relaxation and by coupling to the qubit's . These TLS defects, typically atomic-scale fluctuators, lead to excess loss that scales with the qubit's surface participation. Recent advancements have extended times through material and processing innovations. In 2025, tantalum-based transmon designs achieved T_1 > 200 μs by minimizing surface oxides and losses, with some devices reaching up to 1.68 ms via high-resistivity substrates and optimized encapsulation. As of 2025, further improvements in tantalum-on- designs have achieved times up to 1.6 ms. Surface treatments, such as coatings or etching protocols, have also reduced poisoning—a secondary relaxation channel from non-equilibrium s—thereby enhancing overall T_1 stability. Coherence times are measured using pulsed spectroscopy techniques. T_1 is extracted from exponential fits to the population decay after a \pi-pulse excitation, while T_2 is determined from spin-echo experiments, where a \pi/2-\pi-\pi/2 pulse sequence refocuses low-frequency noise, yielding the echo-limited coherence time T_{2,\text{echo}}. Pure dephasing rates are obtained by fitting Ramsey fringe decay envelopes to Gaussian or exponential models, isolating contributions from 1/f noise spectra.

Noise Reduction

The transmon qubit's design fundamentally addresses charge noise, a dominant decoherence source in earlier superconducting charge qubits, by increasing the ratio of Josephson energy E_J to charging energy E_C (E_J / E_C \gg 1), which exponentially suppresses the charge dispersion \varepsilon(n_g). This suppression, scaling as \varepsilon(n_g) \approx (-1)^n 4 E_C^{n+1/4} e^{- \sqrt{8 E_J / E_C}} for higher levels, reduces sensitivity to offset charge fluctuations n_g by orders of magnitude, enabling robust operation without precise tuning to charge sweet spots. Although primarily insensitive to flux due to its fixed-frequency architecture without an integrated SQUID loop, tunable transmon variants exhibit residual flux noise sensitivity on the order of $10^{-5} to $10^{-6} \Phi_0 / \sqrt{\mathrm{Hz}} at 1 Hz, which is mitigated through fixed-frequency designs that avoid flux-dependent frequency shifts. Dielectric losses from material interfaces and radiation losses via coupling to readout cavities are minimized in transmons through optimized capacitor geometries and large detunings from the cavity mode, yielding Purcell-enhanced decay rates \kappa_P < 1 kHz that do not significantly limit coherence. These noise reduction features collectively enable high-fidelity operations, with single-qubit gate fidelities routinely exceeding 99.9% in modern transmon implementations, compared to below 90% in early Cooper pair box devices plagued by charge noise.

Comparisons

To Charge Qubits

The transmon qubit addresses key limitations of the charge-based (CPB) by minimizing sensitivity to charge noise, a primary decoherence source in early superconducting qubits. In the CPB, qubit energy splitting depends strongly on the dimensionless gate charge offset n_g, necessitating operation at the charge degeneracy point (n_g = 1/2, or "sweet spot") to mitigate first-order charge fluctuations; deviations from this point cause rapid dephasing. The transmon, derived from the CPB but with a large shunt capacitance that increases E_J / E_C to values exceeding 50, suppresses the charge dispersion \varepsilon(n_g) exponentially, rendering it approximately $10^6 times smaller than in the CPB for typical parameters like E_J / E_C = 100. This insensitivity allows transmons to function effectively without precise per-qubit charge tuning, using a global bias voltage instead. This design shift enhances scalability, as CPB devices require individual electrostatic gates and fine-tuned n_g compensation to account for fabrication-induced variations in offset charge, complicating array integration. Transmons, by contrast, operate robustly over a broad range of n_g (e.g., variations of \Delta n_g \approx 0.1 cause negligible frequency shifts), tolerating process imperfections and simplifying control electronics for multi-qubit systems. Performance benefits are evident in coherence metrics: CPB qubits at the sweet spot achieve T_2 times of around 1 μs, constrained by residual charge noise, whereas transmons routinely exceed T_2 > 30 μs, supporting longer quantum gate sequences and easier coupling in arrays. By 2010, these improvements had led major laboratories (e.g., Yale, ) to adopt the transmon as the preferred charge-based , supplanting the CPB due to its superior resilience and fabrication tolerance.

To Other Superconducting Qubits

The transmon offers several advantages over qubits, particularly in times and resilience. qubits, which encode in circulating persistent currents within a superconducting interrupted by Josephson junctions, typically exhibit energy relaxation times (T1) on the order of 10–50 μs, whereas transmons achieve T1 values exceeding 100 μs due to their reduced sensitivity to charge and optimized design in the large charging energy regime. However, transmons rely on for qubit interactions, which can introduce in dense arrays, while qubits enable direct control for tunable coupling, facilitating faster two-qubit gates in some configurations. Compared to phase qubits, which operate as current-biased Josephson junctions with quantum states in the anharmonic potential wells of the tilted washboard, transmons mitigate issues related to escape. Phase qubits are prone to macroscopic quantum tunneling or activation over the potential barrier at finite temperatures, leading to reduced and times below 10 μs in early implementations. In contrast, the transmon's shunted design raises the plasma frequency and enhances (approximately -E_C, where E_C is the charging energy), suppressing such escape processes and enabling gate times below 20 ns with higher operational stability. Hybrid superconducting qubits, such as gatemons, build on the transmon by incorporating nanowires or two-dimensional electron gases as the Josephson weak link, allowing electrostatic gate tunability instead of lines. This parametric amplification approach in gatemons preserves the transmon's charge insensitivity while adding compatibility with fabrication for potential spin-photon interfaces, though it introduces additional disorder from material interfaces that limited early implementations to coherence times of order 1 μs. Subsequent developments have improved coherence times in gatemons to several μs in optimized designs as of 2025. In , transmon-based designs, including , held nearly 62% of the superconducting qubit market share, driven by their adoption in scalable processors by leading firms. Overall, the transmon's insensitivity to charge fluctuations, stemming from the suppression of charge in the E_J >> E_C , facilitates denser two-dimensional planar layouts without isolated islands, supporting toward systems exceeding 1,000 qubits. This contrasts with or designs, which often require more complex three-dimensional wiring or flux lines that hinder planar and increase fabrication challenges for large-scale arrays.

Applications

In Quantum Computing

Transmon qubits form the backbone of scalable quantum processors in , primarily due to their robustness against charge noise and compatibility with planar fabrication techniques. These processors typically employ a two-dimensional , where transmons are arranged in a with nearest-neighbor to enable efficient two-qubit via capacitive or inductive intermediaries. This layout minimizes while supporting the connectivity needed for quantum algorithms, as demonstrated in early multi-qubit devices that scaled from tens to hundreds of qubits. IBM has pioneered large-scale transmon-based processors, with the 2021 Eagle device featuring 127 fixed-frequency transmons in a heavy-hexagonal lattice for improved over prior square grids. Building on this, the Heron r2 advanced to 156 tunable transmons as of 2025, incorporating cross-resonance and enhanced coherence to support deeper circuits for quantum utility demonstrations. In November 2025, IBM announced the with 120 qubits and advanced tunable couplers, aimed at improving and error correction for fault-tolerant computing. These architectures have enabled experiments in quantum and optimization, showcasing transmons' role in pushing toward fault-tolerant computing. In quantum algorithms, particularly surface code error correction, transmons enable the encoding of logical qubits across a large array of physical ones to suppress errors below fault-tolerance thresholds; for instance, a distance-17 code requires over 1,000 physical transmons per logical qubit to achieve logical error rates around 10^{-6} per cycle. By late 2024, IBM demonstrated entanglement of logical qubits using overlapping codes on their processors. These advances highlight transmons' suitability for error-corrected algorithms like quantum transforms and variational quantum eigensolvers. Commercially, integrated transmons into its Bristlecone (72 qubits, 2018) and Sycamore (53 qubits, 2019) processors to achieve quantum advantage in random circuit sampling, with Sycamore's 2D grid demonstrating supremacy over classical supercomputers. In December 2024, Google announced the processor, further advancing transmon-based systems. similarly relies on transmon qubits for its multi-chip systems, such as the 84-qubit Ankaa-2. While focuses on trapped-ion systems, hybrid approaches combining transmon-based superconducting chips with ion traps are under exploration for modular scalability. Transmons dominate superconducting quantum processors, comprising the majority—estimated at over 90%—of deployed multi-qubit chips due to their fabrication maturity and performance. Scaling transmon processors faces challenges from wiring complexity, as increasing qubit counts demand dense interconnects that risk signal loss and thermal loading in cryogenic environments. This has been mitigated through co-design strategies, integrating qubit layouts with multiplexed control electronics and optimized routing algorithms to reduce coaxial lines and enable modular expansion beyond 100 qubits. Such approaches, including tunable couplers for dynamic connectivity, have supported the transition to quantum-centric supercomputing architectures.

As Higher-Dimensional Systems

Transmons, originally designed as two-level qubits, can be extended to function as higher-dimensional quantum systems, or qudits, by utilizing their anharmonic ladder of energy levels beyond the ground and first excited states. This approach leverages the multilevel structure inherent to the transmon's weakly anharmonic potential, where transition frequencies between levels remain sufficiently distinct for selective addressing, enabling encoding of quantum information in dimensions d > 2. For instance, fixed-frequency transmons with high ratios of Josephson energy to charging energy (E_J/E_C up to 325) have demonstrated usability up to d = 12 levels, allowing individual resolution of transitions through microwave pulses. The primary advantage of transmon qudits lies in their ability to access larger Hilbert spaces within a single physical device, potentially reducing the number of elements needed for complex quantum computations compared to arrays of two-level qubits. This compactness facilitates more efficient implementation of multi-qubit operations, such as emulating two-qubit gates via a four-level transmon qudit, which simplifies variational quantum algorithms by lowering circuit depth and resource overhead. Additionally, qudits enable enhanced error correction schemes, where higher-dimensional encoding can suppress logical error rates more effectively than codes, as demonstrated in experiments achieving performance with qutrits (d=3) and ququarts (d=4) encoded in coupled transmon-cavity systems. However, these benefits come with trade-offs, including decreased in higher levels, which increases the risk of leakage errors during operations, and progressively shorter times for excited states due to enhanced coupling to environmental modes. Experimental realizations of transmon qudits have advanced through optimized designs and techniques. In high-E_J/E_C transmons, relaxation times (T_1) range from approximately 64 μs for the lowest levels to 13 μs for the ninth level, with echo coherence times (T_{2E}) approaching twice T_1 under dynamical . Gate fidelities for single-qudit operations remain high, with infidelities below $3 \times 10^{-3} across the lowest 10 levels, and error-per-Clifford metrics as low as $3.25 \times 10^{-4} for 44 ns pulses. Readout fidelity for 10 states has reached 93.8% using multi-tone dispersive measurements combined with classification on existing hardware. Early demonstrations include high-fidelity state transfer between transmon qudits mediated by a , achieving 99.6% fidelity for d=3 and scalability to d=5 with 90.3% fidelity, relying on sequential resonant interactions and to avoid crosstalk. To mitigate measurement challenges, where higher levels exhibit overlapping dispersive shifts in readout resonators, advanced protocols have been developed, such as single-frequency optimization to maximize distinguishability or multi-frequency drives that probe subsets of levels separately. On Quantum hardware, these strategies have improved average readout error rates by adapting drive frequencies beyond qubit-optimized defaults, enabling practical ququart operations like Toffoli gates via two-photon transitions with reduced calibration demands. Overall, transmon qudits represent a pathway to hardware-efficient processing, though ongoing efforts focus on suppressing leakage and extending coherence to higher dimensions for scalable applications.

References

  1. [1]
    An Introduction to the Transmon Qubit for Electromagnetic Engineers
    Jun 21, 2021 · This paper introduces the transmon qubit, a quantum bit used in superconducting circuits, to electromagnetic engineers, aiming to make the ...Missing: definition | Show results with:definition
  2. [2]
  3. [3]
    [cond-mat/0411174] Superconducting Qubits: A Short Review - arXiv
    Nov 7, 2004 · Superconducting qubits are solid state electrical circuits fabricated using techniques borrowed from conventional integrated circuits.
  4. [4]
    A quantum engineer's guide to superconducting qubits
    Jun 17, 2019 · The aim of this review is to provide quantum engineers with an introductory guide to the central concepts and challenges in the rapidly accelerating field of ...From quantum harmonic... · Single-qubit gates · The CPHASE two-qubit gate in...<|separator|>
  5. [5]
    Engineering high-coherence superconducting qubits - Nature
    Sep 23, 2021 · In this Review, the major sources of decoherence in superconducting qubits are identified through an exploration of seminal qubit and resonator experiments.
  6. [6]
  7. [7]
    [0712.3581] Suppressing Charge Noise Decoherence in ... - arXiv
    Dec 20, 2007 · We present an experimental realization of the transmon qubit, an improved superconducting charge qubit derived from the Cooper pair box.
  8. [8]
    [PDF] Quantum process tomography of two-qubit controlled-Z and ...
    We experimentally demonstrate quantum process tomography of controlled-Z and controlled-NOT gates using capacitively coupled superconducting ...
  9. [9]
    IBM's Target: a 4,000-Qubit Processor by 2025 - IEEE Spectrum
    May 10, 2022 · An update to IBM's quantum-computing road map now says the company is on course to build a 4,000-qubit machine by 2025.
  10. [10]
    Princeton puts quantum computing on the fast track with new qubit
    ### Summary of 2025 Tantalum-Silicon Hybrid Transmon Details
  11. [11]
    Charge insensitive qubit design derived from the Cooper pair box
    Feb 28, 2007 · Here, we introduce a new type of superconducting qubit called the "transmon." Unlike the charge qubit, the transmon is designed to operate ...
  12. [12]
    [2103.01225] The superconducting circuit companion - arXiv
    Mar 1, 2021 · This tutorial aims at giving an introductory treatment of the circuit analysis of superconducting qubits, ie, two-level systems in superconducting circuits.
  13. [13]
    Field-based formalism for calculating multiqubit exchange-coupling ...
    (c) Closeup of a transmon formed by a large interdigital capacitor. Here, we ... We can relate the charging energy of the l th transmon to the total shunt ...
  14. [14]
    Cavity-free vacuum-Rabi splitting in circuit quantum acoustodynamics
    Jan 22, 2019 · The IDT capacitance also constitutes the big shunt capacitance of the transmon. Since our model relies on a single parameter to account for both ...
  15. [15]
    Tunable Superconducting Qubits with Flux-Independent Coherence
    Feb 8, 2017 · We have studied the impact of low-frequency magnetic flux noise upon superconducting transmon qubits with various levels of tunability.Missing: seminal SQUID
  16. [16]
    Aluminum Josephson junction microstructure and electrical ... - Nature
    Oct 30, 2024 · Reproducibility of Al/AlOx/Al Josephson junctions is a challenge for scaling up superconducting quantum processors.
  17. [17]
    Surface loss calculations and design of a superconducting transmon ...
    Mar 7, 2022 · The qubit capacitance C ≃ 100 fF is made from two ribbons (shaded gray) of width b − a = 100 μm and length ℓ = 1300 μm, separated by distance 2a ...
  18. [18]
    Advanced CMOS of Superconducting Qubits on 300 mm Wafers
    Sep 18, 2024 · Here we demonstrate superconducting transmon qubits manufactured in a 300 mm complementary metal–oxide–semiconductor (CMOS) pilot line using industrial ...
  19. [19]
    High-performance superconducting quantum processors via laser ...
    May 13, 2022 · We use laser annealing to selectively tune transmon qubits into desired frequency patterns. Statistics over hundreds of annealed qubits demonstrate an ...<|separator|>
  20. [20]
    Optimization of shadow evaporation and oxidation for reproducible ...
    Mar 13, 2023 · Shadow evaporation of Josephson junctions. During electron-beam evaporation of Josephson junctions a diverging metal flow is formed (Fig. 1a). ...
  21. [21]
    [PDF] Perturbative analysis of two-qubit gates on transmon qubits
    In [7] they state that the coupling strength is limited to 10% of the resonator frequency ωr. They report typical values for the coupling strength as g/(2π) ∼ ...<|separator|>
  22. [22]
    Understanding a tuneable coupler circuit — qucat documentation
    By flux biasing the SQUID, we change the effective Josephson energy of the coupler, which modifies the coupling between the two transmons. We will present how ...
  23. [23]
    Tunable coupling of widely separated superconducting qubits
    The resonator coupling strength is g / 2 π = 170 MHz for (a), and g / 2 π = 190 MHz for (b), and the decay rate of the resonator is κ j = ω r j / Q j ⁠. (Here, ...
  24. [24]
    Suppression of crosstalk in superconducting qubits using dynamical ...
    Aug 10, 2021 · Our work demonstrates that dynamical decoupling is an effective and practical way to suppress crosstalk and open system effects.
  25. [25]
    High-fidelity two-qubit gates with transmon qubits using bipolar flux ...
    Sep 5, 2025 · We present a scheme based on superconducting transmon qubits and a control pulse delivery protocol that enables arbitrary controlled-phase gates ...<|control11|><|separator|>
  26. [26]
    [PDF] Mitigation of interfacial dielectric loss in aluminum-on-silicon ...
    The achieved energy relaxation time (T1) of transmon qubits in a planar geometry whose wiring layer is made of aluminum is currently about 100 μs7.Missing: T2 | Show results with:T2
  27. [27]
    [PDF] Investigations of 1/f flux noise in superconducting quantum circuits ...
    It has been shown that low frequency 1/f flux noise is a dominant source of decoherence in phase, flux, and transmon types of qubits [4, 5].
  28. [28]
    Disentangling Losses in Tantalum Superconducting Circuits
    Oct 6, 2023 · Tantalum qubits have recently been demonstrated to exhibit record lifetimes and coherence times exceeding 0.3 ms [15] , which has been ...
  29. [29]
    Wiring surface loss of a superconducting transmon qubit - Nature
    Mar 27, 2024 · A large part of qubits loss is due to microscopic tunneling defects, which form parasitic two-level quantum systems (TLS) and resonantly absorb ...
  30. [30]
    Surface participation and dielectric loss in superconducting qubits
    Oct 19, 2015 · We study the energy relaxation times (T1) of superconducting transmon qubits in 3D cavities as a function of dielectric participation ratios ...
  31. [31]
    2D transmons with lifetimes and coherence times exceeding ... - arXiv
    Mar 19, 2025 · Here we show that replacing the substrate with high-resistivity silicon (Si) dramatically decreases the bulk substrate loss, enabling 2D transmons with time- ...
  32. [32]
    Eliminating Surface Oxides of Superconducting Circuits with Noble ...
    Here, we demonstrate a strategy for avoiding oxide formation by encapsulating the tantalum with noble metals that do not form native oxide.
  33. [33]
    Coherence and Decay of Higher Energy Levels of a ...
    Jan 6, 2015 · We present measurements of coherence and successive decay dynamics of higher energy levels of a superconducting transmon qubit.
  34. [34]
    Charge sensitivity in the transmon regime - arXiv
    Aug 5, 2025 · Koch et al. introduced the transmon qubit in 2007 to mitigate the charge sensitivity inherent to the Cooper pair box design [6] . The transmon ...
  35. [35]
    [2508.16437] Above 99.9% Fidelity Single-Qubit Gates, Two ... - arXiv
    Aug 22, 2025 · Achieving high-fidelity single-qubit gates, two-qubit gates, and qubit readout is critical for building scalable, error-corrected quantum ...
  36. [36]
  37. [37]
    [PDF] How coherent are Josephson junctions? - University of Waterloo
    Coherence in Conventional Transmon Qubit. *. 2. 1. 2. 3.0 s. T. T μ. = = 1. 1.5 s. T μ ... (Ramsey) ~ up to 30 μs, T. 2. (Echo) ~ up to 45 μs. • New observation ...
  38. [38]
    Flux Qubits with Long Coherence Times for Hybrid Quantum Circuits
    Sep 19, 2014 · We now present detailed measurements on the qubits Q3 and Q4 having the longest coherence times. The flux dependence of their frequency, shown ...<|separator|>
  39. [39]
    [PDF] A 4 Superconducting Qubits - Engineering Quantum Systems Group
    Superconducting qubits are solid-state artificial atoms, comprising lithographically defined Joseph- son tunnel junctions and superconducting interconnects [1].
  40. [40]
    Towards practical quantum computers: transmon qubit with a lifetime ...
    Jan 13, 2022 · Here we report a breakthrough in the fabrication of a long lifetime transmon qubit. We use tantalum films as the base superconductor.
  41. [41]
    Semiconductor-Nanowire-Based Superconducting Qubit
    Sep 14, 2015 · For this reason, we refer to our hybrid qubit as a gate tunable transmon, or “gatemon.” Our results highlight the potential of using ...
  42. [42]
    Superconducting Qubit Market Research Report 2033
    Transmon qubits dominate the market, accounting for nearly 62% of ... 14.6 Middle East & Africa Superconducting Qubit Market Size Forecast By Qubit TypeWhat's Driving The Market? · Regional Performance... · Competitor Landscape
  43. [43]
    Charge-insensitive qubit design derived from the Cooper pair box
    Oct 12, 2007 · We introduce a new type of superconducting qubit called the “transmon.” Unlike the charge qubit, the transmon is designed to operate in a regime of ...
  44. [44]
    Material-Driven Optimization of Transmon Qubits for Scalable ... - arXiv
    Aug 7, 2025 · This paper uses design iteration, material analysis, and simulation to optimize transmon qubits, evaluating how materials affect performance.Missing: structure | Show results with:structure
  45. [45]
    Superconducting quantum computers: who is leading the future?
    Aug 19, 2025 · This review examines the state of superconducting quantum technology, with emphasis on qubit design, processor architecture, scalability, and ...
  46. [46]
    Pseudo-2D superconducting quantum computing circuit for the ...
    Apr 21, 2020 · ... 2D, nearest neighbor array of physical qubits. Superconducting quantum circuits have emerged as a major contender as a scalable hardware ...
  47. [47]
    IBM Unveils Breakthrough 127-Qubit Quantum Processor
    Nov 16, 2021 · 'Eagle' is IBM's first quantum processor developed and deployed to contain more than 100 operational and connected qubits.Missing: transmon | Show results with:transmon
  48. [48]
    IBM Debuts Next-Generation Quantum Processor & IBM Quantum ...
    Dec 4, 2023 · IBM Quantum System Two begins operation with three IBM Heron processors, designed to bring quantum-centric supercomputing to reality.
  49. [49]
    IBM Quantum Computers: Evolution, Performance, and Future ...
    Sep 17, 2024 · This paper explores IBM's journey in quantum computing, highlighting key technological achievements, current challenges, and future prospects.Ibm Quantum Computers... · Ii Ibm Quantum · V The Ibm Quantum's Software<|control11|><|separator|>
  50. [50]
    Suppressing quantum errors by scaling a surface code logical qubit
    Feb 22, 2023 · The distance-5 surface code logical qubit is encoded on a 49-qubit subset of the device, with 25 data qubits and 24 measure qubits. Each measure ...Surface Codes With... · Author Information · Google Quantum Ai
  51. [51]
    Demonstration of High-Fidelity Entangled Logical Qubits using ...
    Jul 8, 2025 · We demonstrate the practical utility and advantages of this approach using 127 127 127 127 -qubit IBM quantum processors. Report issue for ...
  52. [52]
    Novera | Rigetti Computing
    A 9-qubit version of Rigetti's most powerful quantum computer. Novera gives you unprecedented access to quantum technology so you can explore further.
  53. [53]
    A scalable routing method for superconducting quantum processor
    Feb 10, 2025 · A well-designed routing scheme can optimize the processor routing spacing, reduce the number of crossovers and corners, lower the influence of ...<|separator|>
  54. [54]
    [PDF] On the Co-Design of Quantum Software and Hardware - NSF PAR
    When fabri- cating a superconducting quantum processor chip, the resonators can only qubits that are physically nearby due to the wire rout- ing constraints.
  55. [55]
  56. [56]
    Systematic study of High $E_J/E_C$ transmon qudits up to $d = 12
    Jul 24, 2024 · In this work, we systematically analyze the trade-offs associated with encoding high-dimensional quantum information in fixed-frequency transmons.
  57. [57]
    Quantum error correction of qudits beyond break-even - Nature
    May 14, 2025 · The transmon hosts a qubit (described by ground and excited states ... Quantum error correction of qudits beyond break-even. Nature 641 ...
  58. [58]
    Improving Transmon Qudit Measurement on IBM Quantum Hardware
    Abstract page for arXiv paper 2307.13504: Improving Transmon Qudit Measurement on IBM Quantum Hardware.
  59. [59]
    Transferring arbitrary d-dimensional quantum states of a ... - Nature
    Aug 1, 2017 · We propose a method to transfer arbitrary d-dimensional quantum states (known or unknown) between two superconducting transmon qudits coupled to a single ...