Fact-checked by Grok 2 weeks ago

Scanning tunneling spectroscopy

Scanning tunneling spectroscopy (STS) is a powerful nanoscale technique that builds upon scanning tunneling microscopy () to probe the local electronic (LDOS) of material surfaces by measuring the quantum tunneling current as a function of applied bias voltage between a sharp metallic tip and the sample. This method enables atomic-scale resolution of electronic properties, distinguishing it from pure STM, which primarily maps surface by maintaining constant current. The foundations of STS trace back to the early 1980s, emerging from the pioneering work of and at , who invented the in 1981 and were awarded the in 1986 for their contributions to nanoscale . STS observations arose during experiments, where variations in tunneling current with voltage revealed spectroscopic features related to surface electronic states, leading to its formal development as a distinct tool for energy-resolved analysis. By the late 1980s, theoretical frameworks, such as those based on Bardeen’s tunneling theory and the Tersoff-Hamann approximation, provided a rigorous basis for interpreting STS data, linking the differential conductance (dI/dV) directly to the LDOS. At its core, STS operates on the principle of elastic tunneling through a barrier approximately 1 nm thick, where the exponentially decaying wavefunctions of s from the and sample overlap, producing a measurable sensitive to both spatial separation and alignment. In practice, the is positioned over the sample using STM feedback for constant height or modes, after which the voltage is ramped (typically from -1 V to +1 V or higher) while recording I-V curves, often averaged over multiple sweeps to reduce . Key operational modes include current imaging tunneling (CITS), which acquires spectroscopic data at every point during a topographic scan, and lock-in amplification for precise dI/dV measurements via small AC modulation of the . Assumptions in data interpretation include a featureless and negligible tip-sample interactions, though advanced models account for many-body effects and voltage-dependent barrier shapes. STS has become indispensable for investigating a wide array of phenomena, including band gaps in semiconductors, superconducting energy gaps, molecular orbitals in organic adsorbates, and quantum confinement in nanostructures. Applications span , such as characterizing topological insulators and 2D materials like , to and chemistry, where it reveals vibrational spectra via inelastic tunneling and enables single-molecule conductance studies. Its high spatial (atomic) and energy (meV) resolution, combined with operability in , low temperature, or electrochemical environments, underscores its role in advancing and .

Overview

Introduction

Scanning tunneling spectroscopy (STS) is a powerful extension of scanning tunneling microscopy () that measures the differential conductance, dI/dV, as a function of applied bias voltage to map the local (LDOS) of a sample's surface at the atomic scale. This technique enables the probing of electronic properties with sub-nanometer , revealing details about energy levels, band structures, and quantum effects in materials. By acquiring spectra at specific locations on the surface, STS provides insights into the electronic landscape that complements the topographic imaging offered by . STS builds directly on the framework by incorporating spectroscopic capabilities through controlled variation of the bias voltage between the sharp metallic tip and the sample, while maintaining a constant tip-sample separation. In contrast to standard , which primarily scans at fixed bias to visualize surface via tunneling current variations, STS captures the energy dependence of the tunneling process. The core principle underlying STS is quantum mechanical tunneling, where the current flows exponentially between the tip and sample without physical contact, and its magnitude reflects the availability of electronic states in both electrodes within the bias energy window. Key advantages of STS include its exceptional atomic-scale for electronic structure mapping, high to local variations in material properties, and adaptability to challenging environments such as (UHV) conditions for clean surface studies or cryogenic temperatures to suppress thermal broadening of spectral features. These attributes make STS invaluable for investigating semiconductors, superconductors, and molecular systems, where it uncovers phenomena like bandgap structures and defect states. The technique was developed in the early 1980s by and at IBM's Zurich Research Laboratory, building on their 1981 invention of , with initial spectroscopic experiments demonstrating rectifying I-V characteristics on gold surfaces in 1982. Their pioneering work on , which laid the foundation for STS, earned them the in 1986, recognizing the transformative impact of tunneling-based .

Historical Development

The invention of the (STM) by and in 1981 at Research Laboratories marked the foundational step toward scanning tunneling spectroscopy (STS), enabling atomic-scale imaging and initial observations of tunneling current variations that hinted at spectroscopic capabilities. Their work, which earned them the 1986 shared with , demonstrated vacuum tunneling between a sharp tip and sample surface, laying the groundwork for measuring electronic properties. Concurrently, theoretical advancements in the early 1980s, notably the 1983 Bardeen-based model by J. Tersoff and D. R. Hamann, established that the tunneling current is proportional to the local density of states (LDOS) near the , providing the conceptual framework for STS to probe energy-dependent electronic structure. Early experimental demonstrations of STS emerged in the mid-1980s, with Robert M. Feenstra and colleagues reporting the first measurements of on cleaved GaAs(110) surfaces in 1986, revealing an apparent band gap of approximately 0.5 due to tip-induced , distinct from the bulk value of 1.42 . This work highlighted STS's potential for characterization, showing sharp onsets in differential conductance corresponding to and conduction band edges. By 1987, Feenstra and Joseph A. Stroscio further refined these measurements, achieving atomically resolved spectroscopic on GaAs(110) and quantifying tip-induced effects. In the 1990s, STS advanced significantly through integration with low-temperature STM systems, enabling detailed studies of . Pioneering low-temperature (4 K) STS on high-Tc cuprates, such as Bi2Sr2CaCu2O8+δ, occurred in the mid-1990s, with significant advancements by J.C. Séamus Davis and collaborators in the late 1990s revealing spatial variations in the superconducting gap, with d-wave symmetry evidenced by nodal states. These experiments, conducted in with superconducting tips, provided direct maps of LDOS modulations linked to pairing mechanisms, transforming understanding of unconventional . Key contributions also came from Heinrich Rohrer's continued refinements in tip engineering and Don Eigler's demonstrations of atomic with functionalized tips in 1990, enhancing spectroscopic resolution. The 2000s saw expansions in STS techniques, including spin-polarized STS (SP-STS) for magnetic imaging, building on Roland Wiesendanger's 1990 room-temperature prototype using ferromagnetic tips on Cr(001). Advancements like antiferromagnetic tips and quantitative spin-resolved spectroscopy enabled atomic-scale mapping of spin textures in materials such as Mn/(110), with spin contrast exceeding 50% in differential conductance. Extensions to non-ultrahigh vacuum environments, including ambient and electrochemical STS, were developed for operando studies, such as on electrodes, allowing real-time LDOS tracking without constraints. Recent milestones up to 2025 include ultrafast time-resolved , with a of lightwave-driven achieving temporal and on defects in WSe2 monolayers, resolving spin-orbit-split states at 50 meV scale. integration has accelerated automated analysis of data for structure discovery on complex surfaces. In 2025, advancements included enhanced for imaging hidden magnetic structures beneath and molecular spin-probe sensing of hydrogen-mediated changes in Co islands. These innovations, alongside contributions from Calvin F. Quate's early work on probe enhancements, continue to propel toward dynamic, high-throughput materials .

Theoretical Foundations

Quantum Tunneling Principles

Quantum tunneling is a quantum mechanical phenomenon in which particles, such as electrons, can traverse regions of space where their classical kinetic energy would be insufficient, effectively penetrating potential barriers that would be impassable under . This occurs because electrons exhibit wave-like behavior, allowing their wavefunctions to extend into and overlap across classically forbidden regions, such as the vacuum gap between a sharp metallic tip and a sample surface in (STM) and (STS). The probability of tunneling arises from the non-zero value of the wavefunction in the barrier region, enabling a measurable current despite the absence of classical conduction paths. In the one-dimensional model of electron tunneling through a , the transmission probability T for an of E incident on a barrier of \phi > E and width d is approximated by T \approx \exp(-2 \kappa d), where \kappa = \sqrt{2m (\phi - E)} / \hbar, with m the and \hbar the reduced Planck's constant. This exponential dependence highlights the sensitivity of tunneling to barrier parameters: small changes in d or \phi lead to dramatic variations in T, as the wavefunction decays rapidly within the barrier. This model, derived from the time-independent for a simple potential step, forms the foundational approximation for understanding transport in junctions like those in /. In the context of STM and STS, the tip-sample separation is typically maintained at approximately 1 , creating a potential barrier of height around 4-5 , corresponding to the average of common metallic tips and samples. This configuration results in an exponentially sensitive tunneling current with respect to the tip-sample distance z, where a change of 0.1 can alter the current by an , enabling atomic-scale resolution. The evanescent wavefunctions in the gap decay as \psi(z) \sim \exp(-\kappa z), with \kappa determined by the barrier height, ensuring that only electrons near the contribute significantly to the overlap between tip and sample states. For reliable STS measurements, coherent elastic tunneling must dominate the process, which requires low bias voltages (<1 V) to minimize inelastic contributions from phonons or other excitations, and low temperatures (typically <10 K) to suppress thermal broadening of the electron distribution and enhance energy resolution. Under these conditions, the tunneling remains phase-coherent, preserving the direct mapping of the sample's local density of states.

Tunneling Current and Density of States

In scanning tunneling spectroscopy (STS), the tunneling current arises from quantum mechanical tunneling of electrons between the tip and the sample, governed by the Bardeen transfer Hamiltonian formalism. This approach treats the tunneling process perturbatively, where the current is determined by the matrix element M_{fi} = \langle \psi_f | H' | \psi_i \rangle, with \psi_i and \psi_f as the initial and final wavefunctions of the isolated tip and sample, respectively, and H' as the interaction Hamiltonian coupling the two systems across the vacuum barrier. The matrix element emphasizes the orbital overlap between tip and sample wavefunctions, which decays exponentially with distance due to the barrier. The tunneling current I(V) at applied bias voltage V is then expressed using Fermi's golden rule as I(V) = \frac{2\pi e}{\hbar} \int_{-\infty}^{\infty} \rho_s(\mathbf{r}, E) \rho_t(\mathbf{r}, E + eV) |M|^2 [f(E) - f(E + eV)] \, dE, where \rho_s(\mathbf{r}, E) and \rho_t(\mathbf{r}, E) are the local densities of states (LDOS) of the sample and tip at position \mathbf{r}, f(E) is the , and |M|^2 is the squared matrix element (often approximated as energy-independent for low biases). This integral accounts for the availability of occupied states in one electrode and unoccupied states in the other, weighted by their LDOS and the tunneling probability. At low temperatures (T \approx 0 K), the Fermi-Dirac distribution becomes a step function, simplifying the expression. Assuming a constant tip LDOS \rho_t(E) (a common approximation for metallic tips), the current for positive bias V > 0 (where the sample is biased positively relative to the ) becomes I(V) \approx \frac{2\pi e}{\hbar} |M|^2 \rho_t \int_{0}^{eV} \rho_s(\mathbf{r}, E) \, dE. This reflects the cumulative contribution from states up to the bias energy. For negative bias V < 0, the roles reverse, probing occupied states below E_F. Positive bias primarily accesses unoccupied sample states above E_F, while negative bias probes occupied states below E_F. The differential conductance dI/dV, obtained by numerically or analytically differentiating the current with respect to voltage, provides a direct measure of the sample's LDOS: \frac{dI}{dV} \propto \rho_s(\mathbf{r}, E_F + eV) \rho_t(\mathbf{r}, E_F), again assuming constant \rho_t and low bias. This proportionality arises because differentiation shifts the energy window, isolating the LDOS at the specific energy E_F + eV. In practice, the Tersoff-Hamann approximation further simplifies this by modeling the tip as an s-wave emitter, where |M|^2 \propto |\psi_s(\mathbf{r}, E_F)|^2, making dI/dV proportional to the sample LDOS at the tip position—ideal for atomic-scale resolution but assuming a featureless tip. Real tips often exhibit d-orbital contributions, which can introduce angular dependence and reduce spatial resolution compared to the s-wave ideal.

Experimental Methods

Instrumentation

Scanning tunneling spectroscopy (STS) relies on the instrumentation of a (STM), which provides the necessary atomic-scale precision and sensitivity to probe local electronic properties. The core components include that enable precise control of the tip position in three dimensions. These scanners, typically constructed from materials like (PZT), achieve sub-angstrom resolution, often better than 0.1 Å in the z-direction and atomic-scale in x-y, allowing for stable raster scanning over sample surfaces. The scanning tip, usually made of or (Pt-Ir) wire, is sharpened to a radius of less than 10 nm via or mechanical cutting to ensure high spatial resolution and minimize artifacts from multiple protruding atoms. The electronic subsystem is critical for controlling and measuring the tunneling process. A bias voltage source applies a direct current (DC) potential between the tip and sample, typically ranging from millivolts to several volts (±10 V), to drive electrons across the vacuum barrier. The tunneling current, on the order of picoamperes (pA), is detected and amplified using a low-noise with sensitivities down to 1 pA or better, often achieving gains of 10^9 V/A. A feedback loop, operating in constant-current mode, adjusts the tip height in real-time via the z-piezo to maintain a setpoint current during topography imaging, ensuring stable tip-sample separation of about 5-10 Å. Environmental controls are essential to minimize contamination and thermal effects that could degrade resolution. Experiments are conducted in ultra-high vacuum (UHV) chambers with base pressures below 10^{-10} Torr to prevent adsorbate layers on the sample surface. Cryogenic cooling stages, often using liquid helium or dilution refrigerators, stabilize the system at temperatures as low as 4 K to reduce thermal drift and broaden spectroscopic features, with thermal stability better than 1 mK. For spectroscopic measurements, a lock-in amplifier is employed to detect the differential conductance dI/dV. This involves superimposing a small alternating current (AC) modulation voltage, typically 1-10 mV at frequencies around 1 kHz, on the DC bias; the resulting AC current component is phase-sensitively demodulated to yield high signal-to-noise ratio spectra proportional to the local density of states. Vibration isolation is paramount for atomic stability, as external mechanical noise can exceed the angstrom-scale tip motion. Systems often feature multi-stage spring suspension platforms combined with eddy current damping using permanent magnets and conductive plates to suppress resonances above 1-10 Hz, achieving isolation better than 40 dB at low frequencies. Recent advancements since 2020 have enhanced STS capabilities through hybrid integrations. Optical access via viewports allows for photo-assisted STS, where laser illumination excites carriers for time-resolved studies of dynamics on femtosecond scales. In 2024, lightwave-driven STS enabled ultrafast probing of atomic-scale defect states with 300 fs resolution. Additionally, combined setups with or enable correlative multimodal characterization of electronic structures in 2D materials and quantum systems, while 2025 developments in facilitate spin-sensitive measurements on single atoms.

Measurement Procedures

Sample preparation is a critical initial step in scanning tunneling spectroscopy (STS) to ensure atomically clean and ordered surfaces under ultra-high vacuum (UHV) conditions, typically below 10^{-10} mbar. For single-crystal semiconductors like silicon, the Si(111)-7×7 reconstructed surface is commonly prepared by mounting the sample on a holder, outgassing at ~900°C to remove contaminants, and then flash-annealing to 1470-1520 K for multiple cycles (e.g., 5 times for 5 s each) followed by controlled cooling to room temperature, yielding large terraces for stable measurements. For layered materials such as highly oriented pyrolytic graphite (HOPG), preparation involves mechanical cleaving with adhesive tape in air or UHV to expose a fresh basal plane, minimizing contamination before transfer to the STM chamber. Thin films or nanostructures are often grown on suitable substrates via techniques like molecular beam epitaxy or sputtering/annealing cycles to achieve epitaxial layers with minimal defects. Once the sample is in position, the metallic tip (typically etched tungsten or platinum-iridium) is approached to the surface in two phases to establish quantum tunneling without contact. Coarse approach uses mechanical mechanisms, such as inertial sliders or micrometers, to position the tip within 1-10 μm of the surface, often under optical or current monitoring. Fine approach then engages to incrementally close the gap, applying a bias voltage (e.g., 0.1-1 V) until a setpoint tunneling current of ~1 nA is reached, confirming a tip-sample separation of ~0.5-1 nm; feedback loops maintain this stability during subsequent operations. STS operates in distinct modes tailored to the sample topography and desired resolution. Point spectroscopy involves parking the tip at a fixed location, disabling the feedback loop, and ramping the bias voltage to record the I(V) characteristic, often with lock-in amplification for dI/dV using a small AC modulation (e.g., 5 mV rms at 1 kHz). For spatial mapping of electronic structure, current imaging tunneling spectroscopy (CITS) acquires full I(V) or dI/dV data at multiple points (typically every pixel) during a topographic scan. Constant-height mode scans the tip over the surface at fixed z-position after initial stabilization, enabling rapid acquisition of spectroscopic data on flat samples like HOPG, though it risks crashes on rough terrains. In contrast, constant-current mode keeps the tunneling current fixed (e.g., 0.1 nA) via active z-feedback during scanning, allowing topography-corrected dI/dV mapping on corrugated surfaces but requiring slower scan speeds to maintain accuracy. Bias ramping in point or mapping spectroscopy employs a linear voltage sweep, commonly from -2 V to +2 V, to probe occupied and unoccupied states symmetrically around the . Each voltage step incorporates a dwell time of ~1-150 ms per point to allow current stabilization and minimize Joule heating or transient effects, with data averaged over multiple sweeps for noise reduction; modulation techniques enhance sensitivity to conductance changes. For spatial resolution of electronic structure, spectroscopic mapping acquires full I(V) or dI/dV spectra on a 2D grid (e.g., 84 × 84 points over 10-100 nm), or measures dI/dV at a fixed bias energy (such as near the , e.g., -0.2 V) while raster-scanning the tip to visualize local density of states () variations. This approach, often in constant-height mode for speed, reveals features like standing waves or band gaps, with the dI/dV signal serving as a direct probe of . Safety protocols are essential to protect the delicate tip and sample from damage. Approach sequences include strict current limits (e.g., compliance at 10-100 nA) to halt motion if tunneling is not detected, preventing uncontrolled crashes into the surface. Piezoelectric actuators exhibit hysteresis (~10-20% nonlinearity), which is calibrated prior to experiments using reference samples or sinusoidal drive corrections to ensure accurate positioning and avoid distortions.

Data Analysis

Spectrum Acquisition

In scanning tunneling spectroscopy (STS), raw data collection begins with disabling the feedback loop to hold the tip-sample separation constant, followed by ramping the bias voltage V across a desired energy range (typically 1-5 V) while recording the tunneling current I as a time-series. This I-V curve captures the electronic response over the voltage sweep, often completed in under 1 minute to limit drift effects. To obtain the differential conductance dI/dV, which is proportional to the local density of states, lock-in detection is employed by superimposing a small AC modulation voltage (e.g., 5-50 mV amplitude at 1-2 kHz frequency) on the DC bias ramp; the in-phase component of the AC current signal, demodulated by the , directly yields dI/dV with high sensitivity. Noise reduction is essential for reliable spectra, as thermal, electronic, and mechanical fluctuations can obscure features. Multiple voltage ramps (typically 10-100 cycles) are averaged at each position to improve the signal-to-noise ratio, with each cycle acquired rapidly to minimize contributions from low-frequency drift. High-frequency noise above 10 kHz is suppressed through the lock-in amplifier's time constant (e.g., 20-50 ms), which acts as a low-pass filter, while additional post-acquisition Gaussian broadening (e.g., 20-40 mV) may be applied to account for instrumental resolution. Preprocessing includes normalization to enhance comparability and isolate intrinsic sample properties. The raw dI/dV is commonly divided by the simultaneously measured I(V) to produce normalized conductance (dI/dV)/I, which mitigates variations due to tip geometry or separation changes and emphasizes relative density-of-states features. For total density of states extraction, dI/dV can be integrated over energy, though this is less common; background subtraction, such as a linear fit to featureless regions, corrects for flat-band shifts or substrate contributions. Spatial mapping extends single-point acquisition to generate energy-resolved images by collecting full I-V or dI/dV spectra on a regular grid, such as 100 × 100 points over a 10 nm² area, with the tip repositioned via closed-loop scanning after each spectrum. These datasets are stitched into hyperspectral volumes, where each spatial pixel contains a full energy spectrum, enabling visualization of local density-of-states variations across the surface. Software tools facilitate real-time control and post-processing; custom LabVIEW programs are widely used for instrument synchronization and immediate feedback during acquisition, while open-source platforms like handle data import, averaging, and visualization of STS curves and maps. Acquisition artifacts arise primarily from instrumental instabilities, including hysteresis due to piezoelectric creep, where the scanner continues to deform after voltage changes, distorting voltage-dependent features in repeated ramps. Thermal drift, causing tip wander at rates up to 1 nm/min at room temperature, is compensated by referencing landmarks (e.g., atomic features) before and after mapping or via post-processing alignment algorithms. Low-temperature operation (e.g., <10 K) and fast ramping further mitigate these issues.

Interpretation Techniques

Interpretation of scanning tunneling spectroscopy (STS) data involves extracting physical properties from differential conductance (dI/dV) spectra, which approximate the local density of states (LDOS) of the sample according to the Tersoff-Hamann approximation. Peaks in dI/dV spectra are assigned based on their energy positions and lineshapes, corresponding to features such as van Hove singularities from saddle points in the band structure, band edges marking the onset of electronic states, or localized impurity states arising from defects or adsorbates. In insulators and semiconductors, symmetric suppression of dI/dV around the indicates band gaps, with the gap width determined by the distance between onset peaks. Fitting models are applied to quantify spectral features. Quasiparticle peaks, often broadened by interactions, are typically fitted with Lorentzian functions to extract peak positions, widths, and amplitudes, revealing lifetimes and coupling strengths. For superconducting samples, the density of states exhibits a gap characterized by coherence peaks at ±Δ, where the gap parameter Δ follows the Bardeen-Cooper-Schrieffer (BCS) theory prediction Δ(0) ≈ 1.76 k_B T_c for weak-coupling superconductors, with spectra fitted to the BCS quasiparticle density of states formula. Deconvolution techniques address instrumental broadening in dI/dV maps. The Tersoff-Hamann model accounts for tip-induced LDOS broadening by convolving the sample LDOS with the tip wavefunction, allowing correction through inverse filtering or simulation-based subtraction. For spatial variations, Fourier analysis of dI/dV maps reveals periodicities from standing waves or moiré patterns, enabling extraction of momentum-space information like band dispersions via Fourier-transform STS. Energy resolution in STS is fundamentally limited by thermal smearing from the Fermi-Dirac distribution, with an effective broadening of approximately 3.5 k_B T (∼1.2 meV at 4 K), and by the modulation voltage amplitude used in lock-in detection, typically set to 1-5 mV to balance resolution and signal-to-noise ratio. Quantitative analysis derives material parameters from spectra. The work function φ can be extracted from the bias threshold where tunneling current onset occurs in I(V) data, calibrated against known tip-sample separations. Band structures inferred from peak positions and dispersions are validated by comparison to density functional theory (DFT) calculations, confirming energy alignments and effective masses, as demonstrated in surface reconstructions like Si(111). Error estimation ensures reliability through statistical methods. Spectra are averaged over multiple pixels in dI/dV maps to reduce noise, with uncertainties derived from standard deviations of fitted parameters, such as band edge positions (±0.03 eV). Confidence intervals are obtained from fit residuals and χ² analysis, quantifying systematic errors from broadening corrections.

Applications

Surface and Material Characterization

Scanning tunneling spectroscopy (STS) plays a crucial role in characterizing the electronic properties of surfaces and materials by providing spatially resolved measurements of the local density of states (LDOS). This technique enables direct probing of band structures, defect states, and surface phenomena at the atomic scale, offering insights into material behavior that complement topographic imaging from (STM). One primary application of STS is the measurement of band gaps in semiconductors, where the tunneling conductance (dI/dV) spectra reveal energy gaps through suppressed current in the forbidden region. For instance, in silicon (Si(111)), STS has visualized a band gap of approximately 1.1 eV, consistent with the indirect bulk value but revealing surface-specific modifications due to reconstruction. Similarly, in gallium arsenide (GaAs), band gaps around 1.4 eV have been mapped, demonstrating how STS distinguishes valence and conduction band onsets with sub-meV energy resolution. STS is particularly effective for studying adsorption and defects, where adsorbates or vacancies induce localized changes in the LDOS. On copper surfaces, carbon monoxide (CO) molecules adsorbed on Cu(100) exhibit prominent LDOS peaks at ~0.5 eV above the Fermi level due to molecular orbitals, allowing identification of bonding configurations. In titanium dioxide (), oxygen vacancies create mid-gap states, observable as sharp LDOS features ~0.8 eV below the conduction band, which influence photocatalytic properties. These measurements highlight how defects can trap charge or alter reactivity. In two-dimensional (2D) materials, STS has elucidated unique electronic structures, such as the linear dispersion of Dirac cones in graphene. On graphene/SiC(0001), dI/dV maps show a V-shaped LDOS profile with a Dirac point at the Fermi level, confirming massless Dirac fermions and enabling bandgap tuning via substrate interactions. For twisted bilayer graphene, STS revealed flat bands and moiré patterns at the "magic angle" of ~1.1° (discovered in 2018), where correlated insulating states emerge due to enhanced electron interactions, as evidenced by suppressed LDOS near the Fermi energy. On metal surfaces, STS identifies surface states and resonances, such as Shockley states on close-packed faces. For copper (Cu(111)), a prominent Shockley surface state appears ~0.4 eV below the Fermi level in dI/dV spectra, forming a parabolic dispersion band that governs surface electron dynamics and scattering. These states are crucial for understanding phenomena like surface diffusion and quantum corrals. Quantitative metrics from STS include work function mapping and Fermi level shifts with doping. Work function variations, derived from apparent barrier height (√(d²I/dz²)/I), have been mapped on heterogeneous surfaces like Si with oxide patches, revealing differences of ~0.5 eV. In doped semiconductors, such as phosphorus-doped Si, Fermi level shifts of up to 0.3 eV toward the conduction band correlate with carrier concentration, providing direct verification of doping efficacy. Integration of STS with STM topography enables correlative analysis for defect classification. By overlaying dI/dV maps on atomic-scale topograms, researchers distinguish vacancy types in materials like , where missing carbon atoms show distinct LDOS depletions compared to Stone-Wales defects, aiding in quality assessment of 2D films.

Advanced Studies in Condensed Matter

Scanning tunneling spectroscopy (STS) has been instrumental in probing the gap symmetry in high-temperature , particularly revealing the pseudogap phase where the density of states exhibits a suppression around the Fermi level without full superconductivity. In underdoped cuprates like , STS measurements show spatial variations in the pseudogap, with d-wave symmetry indicated by asymmetric spectral features that evolve with doping and temperature, distinguishing it from the superconducting gap. These observations support models of competing orders, where the pseudogap arises from phase fluctuations or hidden order parameters. In type-II superconductors, STS enables detailed imaging of vortex cores, where the superconducting order parameter vanishes, leading to localized low-energy states. For instance, in YBa₂Cu₃O₇-δ, vortex cores display subgap conductance peaks attributed to bound quasiparticle states, with fourfold symmetry reflecting the underlying d-wave pairing. Recent studies on overdoped cuprates have uncovered wave-like vortex cores, where extended electronic states propagate along nodal directions, challenging conventional Caroli-de Gennes-Matricon predictions and highlighting quantum interference effects. Such findings, summarized in comprehensive reviews, underscore STS's role in mapping vortex lattice dynamics under magnetic fields. STS has advanced the study of fractional quantum Hall effects by providing high-resolution maps of edge and bulk states in graphene systems. In 2023, experiments on ultra-clean Bernal-stacked bilayer graphene revealed fractional quantum Hall states at filling factors like ν=2/3 and 4/3, with STS spectra showing quantized conductance plateaus and charge density wave signatures at the edges, confirming broken symmetry phases. These measurements, achieving sub-meV energy resolution, demonstrate ideal quantum Hall edge channels confined within nanometers of the physical boundary, free from reconstruction artifacts. In topological materials, STS directly visualizes protected edge states, essential for spintronic applications. On Bi₂Se₃ topological insulators, edge states manifest as linear dispersions crossing the bulk gap, with STS revealing enhanced conductance at step edges due to one-dimensional helical modes robust against backscattering. For Majorana zero modes in hybrid nanowires, such as InSb-Al systems, STS detects zero-bias peaks at wire ends, indicating topological superconductivity, with subgap spectroscopy along the wire confirming particle-hole symmetric excitations localized over tens of nanometers. These signatures, observed in proximitized semiconductors, provide evidence for non-Abelian statistics crucial for quantum computing. Recent methodological advances in STS include ultrafast variants for capturing femtosecond dynamics in quantum systems. In 2024, lightwave-driven STS achieved atomic-scale resolution of phonon-driven charge dynamics on surfaces, revealing transient energy shifts of defect states up to 40 meV with 300 fs temporal resolution through nonlinear tunneling currents. Complementing this, theoretical frameworks for attosecond tunneling microscopy predict coherent electron wavepacket dynamics in nanostructures, enabling real-time observation of strong-field effects. Spin-polarized STS has further elucidated magnetic textures, as in Fe monolayers on W(110), where domain walls exhibit spin-flip excitations and antiferromagnetic coupling, mapped with meV energy and nm spatial resolution. Many-body interactions are probed via STS through signatures like Kondo resonances in magnetic adatoms. On surfaces such as Au(111), single Co or Fe atoms display sharp zero-bias resonances from Kondo screening, with temperature-dependent broadening reflecting the many-body ground state formation. Quasiparticle interference patterns from these adatoms allow band structure mapping, where Fourier-transformed STS reveals scattering vectors modulated by spin-orbit coupling and impurity potentials. From 2023 to 2025, innovations include machine learning integration for automated LDOS analysis in STS datasets. In 2024, k-means clustering sorted large spectroscopic maps to identify electronic orders in iridate materials analogous to cuprates, such as Rh-doped Sr₂IrO₄, enabling unsupervised phase detection with reduced noise and faster processing of terabyte-scale data. Self-supervised denoising techniques further enhanced quasiparticle interference resolution, improving band mapping accuracy by 20-30% in complex materials. For battery interfaces, in-operando STS under electrochemical conditions visualized solid-electrolyte interphase evolution, revealing atomic rearrangements and conductance changes during lithium plating on Cu surfaces.

Limitations

Technical Challenges

One of the primary technical challenges in scanning tunneling spectroscopy (STS) arises from tip preparation, where electrochemical etching of materials like tungsten or platinum-iridium wires often exhibits variability due to factors such as electrolyte concentration, voltage pulses, and immersion time, leading to inconsistent tip sharpness and unintended multi-tip configurations. These multi-tip effects occur when the etching process results in multiple protruding asperities on the tip apex, causing superimposed tunneling currents from several points and distorting spectroscopic signals. Contamination from residual etching solutions or atmospheric exposure further exacerbates artifacts, as adsorbed molecules can alter the tip's electronic properties and introduce noise in current measurements. STS experiments demand ultra-high vacuum (UHV) environments, typically below 10^{-10} mbar, to prevent surface contamination from residual gases that would otherwise adsorb onto the sample or tip, degrading tunneling stability and resolution. At room temperature, thermal drift—arising from differential expansion between the tip, scanner, and sample—poses a significant hurdle, with typical rates exceeding 0.1 nm/s, which can shift the tip-sample gap and compromise data fidelity during extended scans. This sensitivity necessitates cryogenic cooling or advanced stabilization for reliable operation outside controlled UHV setups. Maintaining stability during STS measurements is complicated by piezoelectric creep, where actuators exhibit time-dependent deformation after voltage changes, resulting in position errors up to 10% of the commanded motion and logarithmic drift over minutes. Mechanical vibrations from ambient sources, such as building floors or acoustic noise, further disrupt the tip-sample junction, requiring sophisticated isolation systems to achieve sub-angstrom precision. Additionally, acquiring detailed spectroscopic maps often demands long integration times—frequently hours—to accumulate sufficient current statistics at low tunneling currents, amplifying susceptibility to these instabilities. Sample requirements limit STS applicability, as the technique relies on quantum tunneling that necessitates electrically conductive surfaces with resistivities typically below 1 Ω·cm to ensure stable current flow without charging effects. Insulating materials, such as undoped oxides or wide-bandgap semiconductors, cannot be directly probed and require modifications like doping (e.g., niobium in to induce semiconductivity) or deposition of conductive overlayers to enable measurements. These adaptations, however, can introduce interface complexities that affect intrinsic property characterization. High costs and limited accessibility hinder widespread adoption of STS, with commercial high-end UHV systems exceeding $500,000 due to integrated cryostats, vibration isolation, and vacuum components essential for atomic-scale precision. Operation further demands skilled personnel to handle tip fabrication, vacuum maintenance, and alignment, as misalignment or improper setup can render experiments futile. Extending STS to non-vacuum environments, such as liquids or ambient air, presents scaling challenges, particularly in electrochemical STS where electrolyte solutions serve as the tunneling medium but introduce ionic screening and faradaic currents that limit spatial resolution to tens of nanometers compared to sub-angstrom vacuum performance. These setups suffer from reduced stability due to fluid-induced tip vibrations and potential gradients, restricting applications to specialized in situ studies despite their value for dynamic processes.

Artifacts and Resolution Issues

Scanning tunneling spectroscopy (STS) measurements can be affected by several artifacts that distort the apparent local density of states (LDOS). One prominent artifact is tip-induced band bending, where the strong electric field between the tip and sample surface modifies the potential landscape, particularly in semiconductors or low-density regions, leading to apparent shifts in energy levels. This effect is exacerbated at higher bias voltages and can mimic intrinsic features like band edges. Another common artifact arises from phonon-assisted inelastic tunneling, which introduces steps or peaks in the dI/dV spectra at energies eV > ℏω, where ω is the phonon frequency, due to electron-phonon coupling during tunneling. These features, often observed as sharp onsets in conductance, can be mistaken for electronic transitions if not properly identified. In insulating samples, charging effects further complicate spectra, as trapped charges create local potential variations that suppress or enhance tunneling currents, resulting in asymmetric or suppressed dI/dV signals. Spatial resolution in STS is fundamentally limited by the size of the tip-sample wavefunction overlap, typically achieving lateral resolutions around 0.1 , constrained by the dimensions involved in tunneling. Vertically, the of the tunneling current, I ∝ exp(-κz) with κ ≈ √(2mφ)/ℏ and φ the , enables z-resolutions as fine as 1 pm, allowing detection of subtle height variations. However, energy resolution is finite, broadened by thermal effects (kT, ~25 meV at ), the amplitude of lock-in modulation (~1-10 mV), and lifetime broadening (Γ ~ 1-10 meV in metals). These limits can smear fine spectral features, such as narrow gaps or resonances, reducing the ability to resolve sub-meV structures without cryogenic operation. Common pitfalls include double-barrier effects in thick overlayers, where the additional barrier from the overlayer creates resonant tunneling that modulates the apparent LDOS, often leading to spurious peaks or suppressed transmission in dI/dV maps. In constant-height mode, feedback loop can introduce oscillations or noise in the spectra if the loop is not fully disengaged, coupling mechanical vibrations to the electronic signal. Additionally, dI/dV noise from thermal or instrumental sources can obscure low-energy features, as discussed in spectrum acquisition methods. To mitigate these, high-pass filtering is applied to isolate inelastic features above background elastic tunneling, while simulations, such as calculations, help distinguish true LDOS variations from artifacts by modeling tip-sample interactions. Fundamental limits to resolution in advanced STS variants, such as time-resolved measurements, stem from the Heisenberg uncertainty principle, which trades spatial precision for . Recent 2024 advances in laser-pumped, ultrafast STS have pushed time resolutions to sub-picosecond (300 fs effective via pump-probe), enabling dynamics observation but at the cost of reduced spatial confinement due to broader wavepacket spreading. These trade-offs highlight the intrinsic bounds in probing spatiotemporal electronic processes.

References

  1. [1]
    Assessment of Scanning Tunneling Spectroscopy Modes Inspecting ...
    Mar 18, 2013 · Scanning tunneling spectroscopy (STS) enables the local, energy-resolved investigation of a samples surface density of states (DOS) by ...
  2. [2]
    [PDF] scanning tunneling spectroscopy - Purdue Physics
    Since, in principle, the scanning tunneling spectroscopy experiments probe single molecules on the surface of a gold electrode, it is imperative that the ...
  3. [3]
    September 1981: Invention of the scanning tunneling microscope
    The first researchers to succeed in building an STM were Gerd Binnig and Heinrich Rohrer at IBM Research Laboratories in Zurich, Switzerland, largely because of ...
  4. [4]
  5. [5]
    [PDF] Theory of scanning tunneling spectroscopy - C. Julian Chen
    Julian Chen: Theory of scanning tunneling spectroscopy. ક. ACTING. ATOM-. X. 320 ter (the center of the acting atom) and a direction (the z axis of scanning ...
  6. [6]
    Scanning tunneling spectroscopy of high-temperature ...
    Mar 13, 2007 · Tunneling spectroscopy has played a central role in the experimental verification of the microscopic theory of superconductivity in classical superconductors.Abstract · Article Text · Theory of Electron Tunneling · Low-Temperature Tunneling...
  7. [7]
    Nanoparticle characterization based on STM and STS
    Oct 13, 2014 · In this review, we describe recent progress made in the study of nanoparticles characterized by scanning tunneling microscopy (STM) and ...<|control11|><|separator|>
  8. [8]
  9. [9]
    [PDF] Gerd Binnig and Heinrich Rohrer - Nobel Lecture
    We present here the historic development of Scanning Tunneling Microscopy; the physical and technical aspects have already been covered in a few recent.Missing: seminal | Show results with:seminal
  10. [10]
    Scanning tunneling microscope - IBM
    The scanning tunneling microscope (STM), introduced in 1981 by IBM physicists Gerd Binnig and Heinrich Rohrer, is widely credited with shining a light on ...Missing: spectroscopy seminal paper
  11. [11]
    Ultrafast atomic-scale scanning tunnelling spectroscopy of a single ...
    Mar 14, 2024 · Here we introduce time-resolved lightwave-driven scanning tunnelling spectroscopy (LW-STS) to demonstrate with direct real-space access how ...
  12. [12]
    Automated Structure Discovery for Scanning Tunneling Microscopy
    Apr 21, 2024 · We present automated structure discovery for STM (ASD-STM), a machine learning tool for predicting the atomic structure directly from an STM image.
  13. [13]
    Theory and Application for the Scanning Tunneling Microscope
    Jun 20, 1983 · Theory and Application for the Scanning Tunneling Microscope. J. Tersoff and D. R. Hamann. Bell Laboratories, Murray Hill, New Jersey 07974.Missing: STS | Show results with:STS
  14. [14]
    Work Function: Fundamentals, Measurement, Calculation ...
    Thus, the work functions of most metals are 4 to 5 eV, except for alkali and alkaline earth metals, which generally have work functions around 2 to 3 eV.
  15. [15]
    Theory of the scanning tunneling microscope | Phys. Rev. B
    The theory states tunneling current is proportional to the surface's local density of states at the tip, and lateral resolution is related to tip radius and ...Missing: STS | Show results with:STS
  16. [16]
    A modular ultra-high vacuum millikelvin scanning tunneling ...
    Feb 5, 2020 · We describe the design, construction, and performance of an ultra-high vacuum (UHV) scanning tunneling microscope (STM) capable of imaging at dilution- ...
  17. [17]
    [PDF] Piezoelectric Scanners
    • Piezoelectric scanners are critical elements in SPMs, valued for their sub-angstrom resolution, their compactness, and their high-speed response. However ...
  18. [18]
    Electrochemical Etching of Tungsten for Fabrication of Sub-10-nm ...
    In this study, a two-step etching method is proposed to fabricate a sub-10-nm tungsten tip directly from a 1-mm rod. First, a floating electrolyte-based drop- ...
  19. [19]
    Fast low-noise transimpedance amplifier for scanning tunneling ...
    Jul 1, 2020 · The usual bias voltage range of an STM is ±10 V, which is higher than the permissible supply voltage of the first-stage OP1 (6 V between the ...
  20. [20]
    Circuit design considerations for current preamplifiers for scanning ...
    Apr 12, 2017 · STM preamplifiers typically have transimpedance gains of 109 V/A (1 GΩ) or higher. An ideal preamplifier would also be noise-free with a high ...
  21. [21]
    Cryogen-free modular scanning tunneling microscope operating at 4 ...
    Aug 6, 2024 · We present the design, construction, and performance of a cryogen-free, UHV, low temperature, and high magnetic field system for modular STM operation.
  22. [22]
    Achieving low noise in scanning tunneling spectroscopy
    Oct 2, 2019 · Here, we present a thorough enumeration and analysis of various noise sources and their contributions to the noise floor of STM/S measurements.
  23. [23]
    None
    ### STM Instrumentation Components Summary
  24. [24]
    Analysis of vibration-isolating systems for scanning tunneling ...
    A comparison of one- and two-stage spring-suspended systems with magnetic eddy-current damping shows that one-stage systems suffer from poor isolation at high ...
  25. [25]
    Vibration isolation for scanning tunneling microscopy - AIP Publishing
    We have constructed two types of vibration isolators: a two-stage coil-spring suspension with eddy current dampers ... spring isolator without eddy current damper ...
  26. [26]
    Charge density wave induced nodal lines in LaTe 3 - Nature
    Jun 19, 2023 · The momentum resolution of ARPES represents the error in determining qCDW. The STM measurements were carried out at a base pressure of 2 × 10−11 ...
  27. [27]
    Emergence of Band Structure in a Two-Dimensional Metal–Organic ...
    Jul 17, 2024 · We provide direct evidence of band structure formation upon hierarchical self-assembly, going from metal–organic complexes to a conjugated two-dimensional ...
  28. [28]
    Periodic corner holes on the Si(111)-7×7 surface can trap silver atoms
    May 27, 2022 · Clean and well-ordered Si(111)-7×7 surfaces are prepared by heating the sample substrate by direct current up to 1490 K (5 times for about 5 s) ...
  29. [29]
    Simulated structure and imaging of NTCDI on Si(1 1 1)-7 × 7
    Nov 21, 2014 · Clean Si(1 1 1)-7 × 7 samples were prepared by standard flash annealing to 1200 °C, rapid cooling to 900 °C and then slow cooling to room ...
  30. [30]
    [PDF] LAB UNIT 5: Scanning Tunneling Microscopy
    They will provide you with information about (i) preparation of the experiment, (ii) the procedure for attaining the STM images, (iii) attaining the tunneling ...
  31. [31]
    Assessment of Scanning Tunneling Spectroscopy Modes Inspecting ...
    Mar 18, 2013 · Subsequently, scanning tunneling spectroscopy (STS) was introduced, providing access to the energy dependence of the LDOS and enabling the local ...<|control11|><|separator|>
  32. [32]
    Scanning tunneling co-ramp spectroscopy for reactive adsorbates
    Nov 28, 2011 · The CRS spectra were taken with a sufficiently long dwell time (1 ms) deliberately to cause the decomposition of molecules so that the bias- ...
  33. [33]
    A cryogen-free low temperature scanning tunneling microscope ...
    Jun 1, 2016 · The dwell time per point was 150 ms, and bias modulation was 10 mVrms at 439.7 Hz. FIG. 7. Atomically resolved STM image and tunneling spectrum ...
  34. [34]
    A method to correct hysteresis of scanning probe microscope ...
    Feb 21, 2019 · However, the hysteresis of piezoelectric actuators reduces positioning accuracy and results in distorted SPM images. When regular raster ...
  35. [35]
    (PDF) A method to correct hysteresis of scanning probe microscope ...
    Feb 4, 2019 · However, the hysteresis of piezoelectric actuators reduces positioning accuracy and results in distorted SPM images.
  36. [36]
    Acquisition and analysis of scanning tunneling spectroscopy data ...
    Dec 23, 2020 · Since the development of the scanning tunneling microscope (STM) in 1982,1 spectroscopy measurements with the instrument, i.e., scanning ...
  37. [37]
    Acquisition and analysis of scanning tunneling spectroscopy data ...
    Acquisition and analysis are described for scanning tunneling spectroscopy data acquired from a monolayer of WSe2 grown on epitaxial gra- phene on SiC. Curve ...
  38. [38]
    Scanning tunneling spectroscopy of the surface | Phys. Rev. B
    May 21, 2009 · STS spectra (i.e., d I / d V curves) for the investigation of the sample DOS have been acquired at room temperature, using a lock-in amplifier ...
  39. [39]
    Linewidth of resonances in scanning tunneling spectroscopy
    May 9, 2008 · In most cases, a spectrum corresponds to the average of a few tens of curves allowing a good signal to noise ratio. Figure 1 shows typical ...Missing: multiple | Show results with:multiple
  40. [40]
    Spatially resolved scanning tunneling spectroscopy of single-layer ...
    Sep 8, 2016 · To map the LDOS of the surface, we took I-V curves at points every 0.2 nm along the STS lines as depicted in the STM images in Figs. 3(b)Missing: multiple | Show results with:multiple
  41. [41]
    Gwyddion / Discussion / File formats: Nanonis Spectra - SourceForge
    Apr 18, 2018 · I have been performing some STS measurements using a SPECS Nanonis controller and I am trying to visualise the spectra on a Gwyddion image.
  42. [42]
    Real-space post-processing correction of thermal drift and ...
    Jan 26, 2017 · We have developed a real-space method to correct distortion due to thermal drift and piezoelectric actuator nonlinearities on scanning tunneling microscope ...
  43. [43]
    (PDF) Recent developments in scanning tunneling spectroscopy of ...
    Aug 7, 2025 · In the first case we show that a precise value of the. surface band gap cannot be obtained by simply locating peaks in the normalized ...
  44. [44]
    [PDF] Spectroscopic Evidence for Competing Order-Induced Pseudogap ...
    Abstract The low-energy excitations of cuprate superconductors exhibit various characteristics that differ from those of simple Bogoliubov quasiparticles.
  45. [45]
    Magnetic field–induced pair density wave state in the cuprate vortex ...
    In cuprate superconductors, the vortex cores are surrounded by “halos,” where the density of electronic states exhibits a checkerboard pattern. Edkins et al.
  46. [46]
    Revisiting the vortex-core tunnelling spectroscopy in YBa2Cu3O7−δ
    This cuprate was chosen to revisit the vortex core spectroscopy by STM because the subgap peaks are significantly stronger than analogous low-energy ...
  47. [47]
    Wave Vortex Cores Observed in Heavily Overdoped | Phys. Rev. X
    The objective of the present study is to uncover the true electronic structure of a d -wave vortex core. A first indication of a Wang-MacDonald vortex core [4]
  48. [48]
    Vortex-core spectroscopy of d-wave cuprate high-temperature ...
    Here, we review efforts to image the vortex lattice in copper oxide-based high-temperature superconductors and to measure the characteristic electronic ...
  49. [49]
    [2308.05789] High-Resolution Tunneling Spectroscopy of Fractional ...
    Aug 10, 2023 · We demonstrate high-resolution scanning tunneling microscopy and spectroscopy of fractional quantum Hall states in ultra clean Bernal-stacked bilayer graphene ...
  50. [50]
    Absence of edge reconstruction for quantum Hall edge channels in ...
    Scanning tunneling spectroscopy in graphene reveals ideal quantum Hall edge channels, strongly confined at the physical edge.
  51. [51]
    Characterization of the Edge States in Colloidal Bi2Se3 Platelets
    Apr 16, 2024 · Bulk, three-dimensional Bi2Se3, is a well-known topological insulator (16−19) with a large inverted gap of 200–300 meV and 2D Dirac-cone surface ...Supporting Information · Author Information · Abbreviations · References
  52. [52]
    Subgap spectroscopy along hybrid nanowires by nm-thick tunnel ...
    Oct 20, 2023 · Tunneling spectroscopy is widely used to examine the subgap spectra in semiconductor-superconductor nanostructures when searching for Majorana zero modes (MZMs ...
  53. [53]
    Scanning tunneling spectroscopy of Majorana zero modes in a ...
    Feb 22, 2023 · We describe scanning tunneling spectroscopic signatures of Majorana zero modes (MZMs) in Kitaev spin liquids.Missing: nanowires | Show results with:nanowires
  54. [54]
    Strong-Field Theory of Attosecond Tunneling Microscopy
    Attosecond observations of coherent electron dynamics in molecules and nanostructures can be achieved by combining conventional scanning tunneling ...
  55. [55]
    Spin-polarized scanning tunneling spectroscopy on Fe nanowires
    It is shown that the antiferromagnetic domain structure of the Fe nanowires depends strongly on the miscut of the W (110) substrate. At high miscut the ...
  56. [56]
    Quasiparticle interference from magnetic impurities | Phys. Rev. B
    Jul 13, 2015 · A narrow Kondo resonance is commonly observed around the Fermi energy in these local STS measurements [2, 3] , arising from the formation of a ...
  57. [57]
    Kondo quasiparticle dynamics observed by resonant inelastic x-ray ...
    Oct 17, 2022 · Our results demonstrate how localized electronic degrees of freedom endow correlated metals with new properties.Missing: STS | Show results with:STS
  58. [58]
    Using k-means to sort spectra: electronic order mapping from ... - arXiv
    Aug 13, 2024 · ... 2024. Using k-means to sort spectra: electronic order mapping from scanning tunneling spectroscopy measurements. Report issue for preceding ...
  59. [59]
    Self-supervised learning for denoising quasiparticle interference data
    A. Quasiparticle interference. Scanning tunneling spectroscopy is a powerful tool used to study the electronic properties of materials at the atomic scale.
  60. [60]
    Scanning tunneling microscopy under chemical reaction at solid ...
    Jul 31, 2023 · This review article focuses on recent advances in operando STM, specifically in the study of solid–liquid and solid–gas interfaces.<|control11|><|separator|>
  61. [61]
    [PDF] In-situ Ag tip preparation and validation techniques for scanning ...
    Before querying a sample of interest, it is important for the microscopist to verify the status of the STM tip by conducting STS of a known substance. 1. Image ...
  62. [62]
    Removal of multiple-tip artifacts from scanning tunneling microscope ...
    Nov 14, 2015 · Our analysis clarifies why crystallographic averaging works well in removing the effects of a blunt STM tip (that consists of multiple mini-tips) ...
  63. [63]
    [PDF] Optimization of STM-tip preparation methods ... - DiVA portal
    The aim was to optimize the preparation technique for reproducibility. It has produced several tips with systematically variation of the parameters.
  64. [64]
    An Ultra-High Vacuum Scanning Tunneling Microscope with Pulse ...
    May 30, 2024 · The plates carry the thermal shields, the JT stage and the STM amounting to 17 kg and 3 kg at the 80 K and the 4 K plate, respectively (Fig. 3b) ...
  65. [65]
    [PDF] Thermal Drift Study on SPMs - euspen
    Such imaging process is affected by distortions such as non-linearity of piezoelectric elements, creep, hysteresis, and thermal drift. Software compensation ...<|control11|><|separator|>
  66. [66]
    [PDF] Real-space post-processing correction of thermal drift and ... - arXiv
    We have developed a real-space method to correct distortion due to thermal drift and piezoelectric actuator nonlinearities on scanning tunneling microscope ...
  67. [67]
    [PDF] The Design of a Scanning Tunneling Microscope
    The purpose of the vibration isolator is to eliminate mechanical vibrations that may adversely impact the microscope. The vibration isolator must be capable of ...
  68. [68]
    [PDF] STM / STS of manganese perovskite oxydes - arXiv
    ... STM, which is inoperative on insulators, some CMR manganites remain sufficiently conducting ((ρ ≤ 1Ωcm) for STM experiments to be carried out in the ...
  69. [69]
    Perspectives of cross-sectional scanning tunneling microscopy and ...
    Mar 23, 2018 · In this perspective, we will briefly introduce the basic idea and some recent achievements in using XSTM/S to study complex oxide interfaces.Perspective Article · 3. Cross-Sectional Scanning... · 5. Xstm/s Measurements On...
  70. [70]
    Scanning Tunneling Microscopes - Europhysics News
    in Eu rope, have commercialised complete UHV-. STM systems of this type, designed and engineered to production standards. In spite of their relatively high cost ...<|separator|>
  71. [71]
    Creating a Low-Cost Scanning Tunneling Microscope
    Commercial Scanning Tunneling Microscopes (STMs), which typically cost upwards of $10000, is used in analyzing the atomic structure of conductive materials.Missing: UHV | Show results with:UHV
  72. [72]
    Tunneling Barriers in Electrochemical Scanning Tunneling Microscopy
    A model for the effective barrier height observed with a scanning tunneling microscope. Journal of Electroanalytical Chemistry 1995, 396 (1-2) , 303-307 ...
  73. [73]
    Combining Electrochemical Scanning Tunneling Microscopy with ...
    The overlap of these forces can pose challenges for achieving high-resolution imaging. However, early studies also demonstrated that performing AFM in liquid ...
  74. [74]
    Phonon and plasmon excitation in inelastic electron tunneling ...
    Mar 30, 2004 · We discuss the enhancement of certain phonon modes by phonon-assisted tunneling in STS based on the restrictions imposed by the electronic structure of ...
  75. [75]
    Feedback loop dependent charge density wave imaging by ...
    The impact of the feedback loop setting on STS imaging has been the subject of previous investigations, all suggesting that scanning the tip in constant height ...
  76. [76]
    Scanning Tunneling Spectroscopy of Molecules on Insulating Films
    Aug 9, 2025 · Second, the ionic relaxations in a polar insulator may lead to an interesting charge bistability in atoms and molecules. STM-based molecular ...