X-ray fluorescence (XRF) is a non-destructive analytical technique that determines the elemental composition of materials by exciting atoms in a sample with high-energy X-rays, causing the emission of characteristic fluorescent X-rays that are detected and analyzed to identify and quantify elements from sodium to uranium.[1] The method is widely used for routine chemical analysis of solids, liquids, and powders in fields such as geology, environmental monitoring, and industrial quality control due to its speed, minimal sample preparation, and ability to detect major, minor, and trace elements at concentrations as low as 1 ppm.[1][2]The fundamental principle of XRF is rooted in atomic physics: when a sample is irradiated with primary X-rays of sufficient energy (typically from an X-ray tube operating at 15–50 kV), photons interact with inner-shell electrons, ejecting them and creating vacancies.[3][2] Outer-shell electrons then transition to fill these vacancies, releasing secondary X-rays with discrete energies characteristic of the element involved, such as the Kα line at approximately 25.27 keV for tin or 10.53 keV for arsenic.[4][2] These emissions follow Moseley's law, which relates the square root of the X-ray frequency to the atomic number, enabling precise elemental identification.[3]Background radiation, including bremsstrahlung (continuous spectrum from electron deceleration) and scattering (Rayleigh or Compton), can interfere but is managed through filters or spectral processing.[4][5]XRF instrumentation typically employs either energy-dispersive (EDXRF) or wavelength-dispersive (WDXRF) spectrometers. EDXRF uses semiconductor detectors, such as silicon drift detectors, to measure photon energies directly, offering portability for field applications like portable XRF (pXRF) units weighing under 2 kg, though limited to elements with atomic numbers Z ≥ 11–20 and prone to spectral overlaps for low-Z elements.[4][2] In contrast, WDXRF achieves higher resolution via crystal diffraction (Bragg's law), allowing detection of lighter elements (Z = 11–25) and simultaneous multi-element analysis in sequential or fixed configurations, ideal for laboratory precision in industries like cement and metallurgy.[1][4] Quantitative analysis requires calibration with standards, as intensity depends on factors like matrix effects and self-absorption.[5]Historically, XRF builds on Wilhelm Conrad Röentgen's 1895 discovery of X-rays, with the technique's foundations laid in 1913 by Henry Moseley, who demonstrated the characteristic X-ray spectra's relation to atomic number.[5][3] Practical qualitative and quantitative applications emerged in 1923 through the work of George de Hevesy and Dirk Coster, who used XRF to discover the element hafnium, establishing XRF as a versatile tool for non-destructive analysis across scientific and industrial domains.[5][6] As of 2025, advancements such as deep learning-based signal extraction and new macro-XRF scanners continue to enhance sensitivity for trace-level detection in environmental, forensic, pharmaceutical, and cultural heritage contexts.[7][8]
Fundamentals
Characteristic Radiation
Characteristic radiation refers to the discrete X-ray emissions produced when an atom is ionized by the ejection of an inner-shell electron, creating a vacancy that is subsequently filled by an electron from a higher energy shell, with the energy difference released as an X-ray photon whose energy is unique to the element's atomic structure.[9] This process occurs in X-ray fluorescence following excitation by higher-energy radiation, where the emitted X-rays serve as a fingerprint for elementalidentification./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/08%3A_Atomic_Structure/8.06%3A_Atomic_Spectra_and_X-rays)The characteristic X-rays are classified based on the electron shells involved in the transition, denoted as K, L, and M shells corresponding to principal quantum numbers n=1, 2, and 3, respectively. K-shell emissions, the most energetic and commonly used in analysis, arise from transitions to the K shell: Kα lines result from L-shell electrons (n=2) filling the K vacancy, with Kα1 and Kα2 denoting specific sub-shell transitions (L3 to K and L2 to K), while Kβ lines involve M-shell electrons (n=3) filling the vacancy, such as Kβ1 from M3 to K. L-shell emissions (Lα, Lβ) occur when vacancies in the L shell are filled by M or higher electrons, and M-shell lines are weaker and typically observed in heavier elements. These lines have fixed energies determined by the binding energies of the shells, which increase with atomic number.[10][9]The relationship between the frequency of these characteristic X-rays and the atomic number Z is described by Moseley's law, an empirical relation derived from experimental measurements. The law states that the square root of the X-ray frequency ν is linearly proportional to (Z - b), where b is a screening constant (approximately 1 for K lines), expressed as:\sqrt{\nu} = c (Z - b)Here, c is a constant related to the Rydberg frequency adjusted for relativistic effects in inner shells. This law, established through systematic spectroscopy of elements, confirmed the atomic number as a fundamental property and predicted missing elements in the periodic table.[11]Binding energies for inner shells vary by element; for example, the K-shell binding energy of iron (Z=26) is approximately 7.11 keV, leading to a Kα emission at 6.4 keV when an L-shell electron transitions to fill the vacancy, with the energy difference reflecting the L-shell binding of about 0.7 keV. Similar patterns hold for other elements, such as copper (Z=29) with Kα at 8.0 keV, enabling precise elemental mapping in fluorescence spectra.[12][13]In contrast to characteristic radiation, bremsstrahlung (braking radiation) produces a continuous spectrum of X-rays from the deceleration of charged particles, such as electrons in an X-ray tube, forming a broad background that underlies the sharp characteristic peaks but lacks elemental specificity.[14] This continuum must be subtracted or accounted for in spectral analysis to isolate the discrete characteristic lines.
X-ray Interactions with Matter
In X-ray fluorescence (XRF), the dominant interaction mechanism for exciting atoms in a sample is the photoelectric effect, whereby an incident X-rayphoton transfers its energy to an inner-shell electron, ejecting it if the photon energy exceeds the electron's binding energy and creating a vacancy.[15] This process is particularly efficient in the energy range typical for XRF (approximately 1–50 keV), where the photoelectric absorption cross-section \sigma scales approximately as \sigma \propto Z^4 / E^3, with Z being the atomic number of the absorbing atom and E the photon energy.[16] The strong dependence on Z favors excitation in higher-Z elements, while the inverse cubic dependence on E ensures dominance over other interactions at lower energies.Secondary interactions include Compton scattering (incoherent scattering), which imparts partial energy transfer to an electron, resulting in a wavelength shift and contributing to a broad continuum background in XRF spectra, and Rayleigh scattering (coherent scattering), which elastically scatters photons without energy loss, producing narrow peaks that add to spectral noise.[15] These scattering processes do not generate characteristic fluorescence but increase the overall background, reducing signal-to-noise ratios, especially in lighter matrices where Compton effects are more pronounced.[15]Absorption edges, such as the K-edge (for the 1s orbital) and L-edges (for 2s and 2p orbitals), manifest as sharp discontinuities in the absorption cross-section when the incident photon energy matches the binding energy of the respective shell electrons. Above an absorption edge, the photoelectric cross-section increases abruptly, enabling inner-shell ionization and subsequent fluorescence; however, the efficiency of X-ray emission versus competing non-radiative processes (like Auger electron emission) is governed by the fluorescence yield \omega_K, which for mid-Z elements (Z ≈ 20–50, e.g., calcium to tin) ranges from approximately 0.16 to 0.86, typically 0.3–0.9 for many analytical cases.[17]Matrix effects arise from the sample's composition influencing both the attenuation of incident (primary) X-rays and the escape of fluorescent X-rays, primarily through photoelectric absorption by surrounding atoms.[18] These effects are quantified by the linear absorptioncoefficient \mu (in cm⁻¹), which describes exponential attenuation via the Beer-Lambert law I = I_0 e^{-\mu x}, where x is path length; heavier matrix elements enhance absorption of lower-energy fluorescence, while lighter ones may cause secondary excitation (enhancement).[18]The effective depth of XRF analysis is limited by the penetration depth of primary X-rays into the sample (typically 10–100 μm for keV energies in solids) and the escape depth of fluorescent X-rays (often 1–10 μm, shorter for lower energies due to higher absorption).[15] These depths vary with photon energy, sample density, and composition, confining XRF to surface or near-surface analysis in most solid matrices unless using low-density or thin samples.[15]
Excitation Sources
X-ray fluorescence (XRF) relies on primary radiation to excite atoms in a sample, ejecting inner-shell electrons and prompting the emission of characteristic fluorescent X-rays. The excitation sources generate this primary radiation through various mechanisms, primarily leveraging photoelectric absorption to ionize target atoms.[19] Common sources include X-ray tubes, radioisotopes, synchrotron radiation, and secondary fluorescers, each offering distinct advantages in intensity, energy range, and portability for analytical applications.[20]X-ray tubes are the most widely used excitation sources in laboratory XRF systems, producing a spectrum of bremsstrahlung continuum radiation alongside characteristic lines from the anode material.[20] In these tubes, high-voltage electrons from a heated cathodefilament strike the anode, generating bremsstrahlung when electrons decelerate and characteristic X-rays when they eject inner-shell electrons from anode atoms.[19]Anode materials such as tungsten, rhodium, or molybdenum are selected to match the excitation needs; for instance, tungsten provides a broad bremsstrahlung spectrum up to the tube's maximum energy, while rhodium yields prominent K-lines around 20 keV for mid-range elements.[21] Tube operation typically involves voltages of 20-60 kV and currents adjusted from milliamperes to amperes to control intensity and penetration depth, enabling excitation of elements from sodium to uranium depending on the setup.[22] The invention of the Coolidge hot-cathode X-ray tube in 1913 marked a pivotal advancement, providing a stable, controllable source that replaced unreliable gas tubes and enabled the practical development of XRF spectroscopy, as demonstrated in early elementalidentification experiments.[23]Radioisotope sources offer compact, battery-independent excitation for portable XRF analyzers, emitting monoenergetic X-rays or gamma rays via radioactive decay.[24] Common examples include iron-55 (^55Fe), which decays to produce 5.9 keV Mn Kα X-rays suitable for light elements like sodium to calcium, and americium-241 (^241Am), emitting 59.5 keV gamma rays for exciting higher atomic number elements up to uranium.[25] These sources are sealed in low-activity capsules (typically 0.1-1 Ci) to ensure safety and longevity, with half-lives of about 2.7 years for ^55Fe and 432 years for ^241Am, making them ideal for field-deployable instruments despite lower intensity compared to tubes.[20]Synchrotron radiation serves as a premium excitation source in advanced research facilities, delivering exceptionally high-intensity, tunable beams for micro- and nano-scale XRF.[26] Produced by accelerating charged particles in storage rings, synchrotron sources generate brilliant X-ray beams with fluxes orders of magnitude higher than laboratory tubes, enabling trace element detection at parts-per-billion levels.[27] Their tunability across a wide energy range (from soft to hard X-rays) allows precise selection of excitation energies to minimize background and enhance sensitivity for specific elements, while natural linear polarization reduces scattering artifacts in anisotropic samples.[28] Although limited to synchrotron labs due to infrastructure demands, these sources excel in applications requiring spatial resolution below 1 μm.[29]Secondary fluorescers enhance excitation specificity by converting broad-spectrum primary radiation into targeted monoenergetic beams, often using filters or dedicated targets. In this approach, an initial X-ray tube or source irradiates a secondary target material, such as molybdenum for ~17.5 keV Kα excitation, which then fluoresces characteristic X-rays directed at the sample.[30] Filters, like aluminum or zirconium foils, are paired with these targets to suppress unwanted bremsstrahlung and lower-energy lines, optimizing the beam for elements in specific atomic number ranges while reducing background noise by up to 100-fold compared to direct excitation. This method is particularly valuable in energy-dispersive XRF for improving signal-to-noise ratios without requiring high-power sources.[3]
Detection and Measurement
Dispersion Methods
In X-ray fluorescence (XRF) analysis, dispersion methods are essential for separating the characteristic X-rays emitted from a sample according to their wavelengths or energies, thereby enhancing spectral resolution and enabling accurate identification of elemental lines. These techniques distinguish overlapping spectral features that would otherwise compromise quantification, particularly in complex matrices with multiple elements. The two primary approaches are wavelength-dispersive (WD) and energy-dispersive (ED) methods, each offering trade-offs in resolution, speed, and instrumental complexity.Wavelength-dispersive XRF (WDXRF) employs diffraction from analyzing crystals to achieve separation, governed by Bragg's law, which describes the condition for constructive interference of X-rays reflected from crystal lattice planes. The law is expressed as n \lambda = 2 d \sin \theta, where n is the order of diffraction (typically 1 for first-order spectra), \lambda is the wavelength of the X-ray, d is the interplanar spacing of the crystal lattice, and \theta is the angle of incidence relative to the lattice planes.[31] This diffraction selectively reflects X-rays of a specific wavelength at a given angle, allowing isolation of individual characteristic lines. In practice, WDXRF systems use goniometers—either fixed for simultaneous multi-element analysis or scanning for sequential measurement—to position the analyzing crystal and detector such that the Bragg condition is satisfied for the desired wavelength.[32] Fixed goniometers employ multiple crystals in parallel channels for rapid, simultaneous detection of several elements, while scanning designs vary \theta to sweep through the spectrum.[33]In contrast, energy-dispersive XRF (EDXRF) avoids physical dispersion altogether, instead relying on electronic pulse-height analysis to sort photons by their energy. Here, the full spectrum of fluorescent X-rays is collected by a detector that generates voltage pulses proportional to the incident photon energy, with a multichannel analyzer sorting these pulses into an energy histogram. This approach provides a non-dispersive separation based on the direct relationship between X-ray energy E and wavelength \lambda via E = hc / \lambda, where h is Planck's constant and c is the speed of light, allowing rapid acquisition of the entire spectrum without mechanical scanning.[34]To improve beam focusing and collection efficiency in both WD and ED systems, specialized optics such as multilayer mirrors and polycapillary optics are employed. Multilayer optics consist of alternating thin films (e.g., heavy metal layers like tungsten separated by light spacers like carbon) that act as synthetic Bragg reflectors, enabling efficient reflection and focusing of X-rays over a broad energy range while mimicking large d-spacings for low-angle diffraction.[35] Polycapillary optics, formed from bundles of fine glass capillaries, guide X-rays through successive total internal reflections, collecting a large solid angle from the source and focusing it to a small spot (typically 10-100 μm) on the sample with minimal loss, which is particularly useful for micro-XRF applications.[36]The resolution of these methods differs significantly, impacting their suitability for trace analysis and peak deconvolution. EDXRF typically achieves an energy resolution of about 150 eVfull width at half maximum (FWHM) at 5.9 keV (manganese Kα line), limited by detector electronics and noise, which can lead to overlaps in closely spaced lines. WDXRF, however, provides superior resolution of 5-20 eV FWHM, thanks to the precise angular selection via crystal diffraction, enabling clear separation of closely spaced lines such as the Kα line of arsenic (~10.51 keV) and the Lα line of lead (~10.55 keV) (ΔE ≈ 40 eV).[13]
Detector Technologies
In X-ray fluorescence (XRF) spectrometry, detectors convert incident fluorescent X-rays into electrical signals for analysis, with performance determined by factors such as energy resolution, count rate capability, and sensitivity to specific energy ranges. Gas-filled detectors, particularly proportional counters, have been widely used since the early development of XRF systems due to their robustness and suitability for moderate-resolution applications. These detectors operate by ionizing a gas mixture within a chamber, where the resulting electron cloud is amplified proportionally to the incident X-ray energy. A common filling gas is P10, consisting of 90% argon and 10% methane, which provides stable operation at atmospheric pressure and flow rates around 25 sccm.[37] Typical energy resolution for these counters is approximately 20% full width at half maximum (FWHM) at 5.9 keV, enabling elemental identification but limiting overlap resolution for closely spaced lines.[38]Solid-state detectors, such as lithium-drifted silicon [Si(Li)] devices, offer significantly improved energy resolution compared to gas-filled types, making them a cornerstone for high-precision XRF measurements. In Si(Li) detectors, X-rays generate electron-hole pairs in the depleted silicon layer, with the charge collected and proportional to the photon energy. These detectors achieve resolutions of 130-180 eV FWHM at 5.9 keV, allowing better separation of characteristic lines from light to medium elements.[39] To minimize thermal noise and leakage current, Si(Li) detectors require cooling to liquid nitrogen temperatures, approximately -196°C.[40] This cryogenic operation, while effective, necessitates bulky cryostats, which can complicate portable or field-based XRF setups.Silicon drift detectors (SDDs) represent an advancement over traditional solid-state detectors, combining high resolution with practical operational advantages for modern XRF instruments. SDDs employ a radial electric field to drift charge carriers to a small anode, reducing capacitance and electronic noise independent of active area size. They operate at room temperature or with minimal Peltier cooling, enabling resolutions around 120 eV FWHM at 5.9 keV and supporting high count rates exceeding 10^5 counts per second (cps) without significant degradation.[41] This capability is particularly valuable in energy-dispersive XRF (EDXRF) for analyzing samples with intense fluorescence signals.Scintillation detectors, typically using thallium-doped sodium iodide [NaI(Tl)] crystals coupled to photomultiplier tubes (PMTs), are employed in wavelength-dispersive XRF (WDXRF) systems for detecting higher-energy X-rays from heavier elements. In these devices, incident X-rays excite the scintillator to produce visible light photons, which are amplified by the PMT to generate a measurable electrical pulse. NaI(Tl) scintillators provide good efficiency for energies above 10 keV but coarser resolution, around 8-10% FWHM at 662 keV, making them complementary to finer-resolution detectors in multi-channel WDXRF setups.[42]In high-flux environments, such as those encountered with intense excitation sources or concentrated samples, detector performance is limited by dead time—the interval after an event during which the detector cannot register another—and pile-up effects, where overlapping pulses distort energy measurements. Dead time arises from the finite processing duration of each signal, leading to count rate losses that can exceed 50% at fluxes above 10^4 cps in older systems.[43] Pile-up occurs when multiple X-rays arrive within this dead time, summing their energies and broadening peaks or creating false high-energy events, which necessitates corrections to maintain quantitative accuracy.[44] Modern detectors like SDDs mitigate these issues through faster readout electronics and advanced pile-up rejection algorithms.
Signal Processing and Intensity Analysis
In X-ray fluorescence (XRF) spectrometry, signal processing begins with the amplification and shaping of electrical pulses generated by the detector in response to incoming X-ray photons. Pre-amplifiers, typically charge-sensitive types, convert the initial charge from the detector into a voltage step whose amplitude is proportional to the photon's energy, while providing initial noise filtering.[45] These signals are then fed into main amplifiers, which apply pulse shaping—often using CR-RC or Gaussian filters—to optimize signal-to-noise ratio and energy resolution by narrowing the pulse width and reducing ballistic deficits. Shaping times are commonly set between 1 and 10 μs, balancing resolution (better at longer times) against count rate limitations (higher at shorter times).[46]The shaped pulses undergo pulse-height analysis to construct the energy spectrum. An analog-to-digital converter (ADC), usually with 10-14 bit resolution and sampling rates exceeding 10 MHz, digitizes the peak amplitude of each pulse, assigning it to a corresponding channel in a multichannel analyzer (MCA). This histogram of channel counts versus energy forms the raw XRF spectrum, where characteristic peaks correspond to elementalemission lines.[43] Modern digital pulse processors often integrate ADC and MCA functions, enabling real-time pile-up rejection and improved throughput for high-flux applications.[47]Intensity analysis quantifies the net X-ray counts under characteristic peaks to determine elemental concentrations. The fluorescence intensity I for an element is fundamentally given by I = k \cdot N \cdot \omega \cdot (1/\mu), where k is an instrumental constant, N is the atom density of the element, \omega is the fluorescence yield, and \mu is the mass absorption coefficient accounting for self-absorption in the sample. Peak intensities are measured as the integrated area above the baseline, often fitted using Gaussian or Voigt profiles to account for instrumental broadening.Background subtraction is essential to isolate true fluorescence signals from continuum contributions like Bremsstrahlung (from the excitation source) and Compton (inelastic scattering) tails, which can obscure low-concentration peaks. Common methods include the sensitive nonlinear iterative peak-clipping (SNIP) algorithm, which iteratively estimates and subtracts the background by clipping peak regions while preserving continuum shape, and Fourier transform techniques that filter high-frequency spectral components.[48] For Compton tails specifically, empirical modeling or subtraction using spectra from pure scatterers (e.g., carbon) corrects asymmetric broadening from detector response.[8]Quantitative corrections address matrix effects, where absorption and secondary enhancement alter measured intensities. The fundamental parameters (FP) method employs Sherman's equation to iteratively compute concentrations by modeling excitation, yield, and absorption physics without standards, ideal for unknown samples. Empirical calibration, conversely, uses reference standards to derive influence coefficients (e.g., via Lachance-Traill alpha factors) that adjust for inter-element interactions, offering simplicity for routine analysis of similar matrices. Both approaches ensure accuracy within 1-5% for major elements, depending on matrix complexity.
Analytical Methods
Energy-Dispersive X-ray Fluorescence
Energy-dispersive X-ray fluorescence (EDXRF) systems employ a direct irradiation approach where the sample is excited by a polychromatic X-ray beam from a tube source, without the use of a monochromator or dispersing element. The emitted fluorescent X-rays are collected by an energy-sensitive detector, such as a silicon drift detector, which measures the energy of incoming photons and generates a full spectrum across a wide energyrange simultaneously. This architecture allows for the acquisition of the entire spectrum in seconds, enabling rapid multi-element analysis from sodium to uranium.[49][50][51]The primary advantages of EDXRF include its high speed, with analyses typically completed in seconds to minutes, making it suitable for high-throughput screening. Its simpler design, lacking moving parts or complex optics, facilitates portability, with handheld units emerging in the 1990s for field use, and overall cost-effectiveness due to efficient source utilization and minimal maintenance. These features make EDXRF ideal for on-site applications where rapid, non-destructive elemental mapping is required.[52][49][50]However, EDXRF systems suffer from lower energy resolution, typically 150–300 eV, compared to other methods, which can lead to peak overlaps for closely spaced elements and artifacts like escape peaks in silicon-based detectors where a fraction of the photon energy is lost to siliconfluorescence. To mitigate scatter from the excitation source and background noise, EDXRF setups often incorporate filters or attenuators to selectively absorb low-energy radiation, as well as secondary targets in Cartesian geometry configurations that fluoresce at specific energies while suppressing direct beam scatter.[51][50][30]In practical applications, such as miningexploration and alloy composition analysis, EDXRF achieves detection limits of approximately 10–150 ppm for mid-Z elements like iron and copper under typical conditions (e.g., 180 s acquisition time), enabling efficient sorting of ores and quality control of metal alloys without sample preparation. For instance, portable EDXRF units are widely used in the field to assess ore grades in real-time, supporting decisions in resource extraction.[53][54]
Wavelength-Dispersive X-ray Fluorescence
Wavelength-dispersive X-ray fluorescence (WDXRF) spectrometry employs a sequential scanning mechanism to achieve high spectral resolution, enabling precise measurement of individual elemental emission lines in complex samples. In this configuration, the system utilizes a goniometer that positions a primary analyzing crystal and an associated detector to diffract and detect X-rays at specific Bragg angles corresponding to desired wavelengths, allowing sequential analysis of multiple elements by mechanically scanning through the spectrum.[55][56]The core of the WDXRF system involves monochromators, which can be flat or curved crystals designed to select and focus specific wavelengths from the fluorescent X-rays emitted by the sample. Curved crystals, often configured in Johansson geometry, provide exact focusing by bending the crystal to a radius equal to the Rowland circle diameter, optimizing intensity and resolution for a range of wavelengths; this geometry is particularly effective for maintaining a narrow bandpass and reducing aberrations in the diffracted beam. Flat crystals, while simpler, are used in certain setups for broader dispersion but with less focusing efficiency.[57][58][59]Analyzing crystals are selected based on the target element's emission energy to maximize diffraction efficiency via Bragg's law. For high-energy lines (typically >4 keV, such as those from iron or heavier elements), lithium fluoride (LiF, often the 200 plane with d-spacing of 0.2014 nm) is commonly used due to its suitability for short wavelengths and high resolution. In contrast, for low-energy lines from light elements (<2 keV, like sodium or magnesium), pentaerythritol (PET, d-spacing ~8.74 Å) is preferred as it accommodates longer wavelengths with better sensitivity.[60][61][62]Detection in WDXRF systems relies on specialized counters tailored to energy ranges: flow proportional counters, filled with gases like argon-methane, are employed for low-energy X-rays (light elements) due to their high efficiency in registering soft radiation through ionization; scintillation counters, using materials like NaI(Tl), are used for high-energy X-rays (heavy elements) for their superior counting rates and energy discrimination. This dual-detector approach enhances overall sensitivity across the spectrum.[63][64]The primary advantages of WDXRF include its exceptional spectral resolution, which minimizes overlaps and interferences between emission lines, enabling accurate quantification even in matrices with closely spaced peaks. This results in low detection limits, often below 1 ppm for trace elements, making it suitable for demanding analyses. WDXRF is widely applied in industries such as cement production for oxide composition control and steel manufacturing for alloying element monitoring, where precision and reliability are critical for quality assurance.[32][51][65][32][66]
Applications and Extensions
Elemental and Chemical Analysis
X-ray fluorescence (XRF) serves as a primary technique for elemental and chemical analysis, enabling the identification and quantification of elements in diverse materials without sample destruction. This non-destructive method is widely applied across industries to determine bulk compositions, detect trace contaminants, and ensure regulatory compliance, offering rapid results that support quality control and environmental monitoring.[67]Quantitative analysis in XRF relies on establishing relationships between measured X-ray intensities and element concentrations, typically through empirical calibration curves or the fundamental parameters (FP) method. Calibration curves are generated using standards with known concentrations, plotting XRF peak intensity against concentration to derive linear or polynomial fits for accurate quantification.[67] The FP method, in contrast, models the physical processes of X-ray excitation and emission theoretically, requiring minimal standards and accounting for matrix effects to enable analysis from sub-parts-per-million (ppm) levels up to 100% concentrations.[68] This versatility allows XRF to quantify major, minor, and trace elements simultaneously, with detection limits often reaching 1-10 ppm for many metals depending on the instrument.[67]In the battery industry, XRF is extensively used for non-destructive elemental analysis along the value chain, including raw material quality control, cathode and anode composition verification, and black mass recycling from end-of-life lithium-ion batteries to recover valuable metals like lithium, cobalt, and nickel. As of 2025, portable XRF analyzers facilitate on-site sorting and process monitoring in recycling facilities, supporting the global transition to sustainable energy storage.[69][70]Detecting light elements (atomic number Z ≤ 12, such as sodium (Z=11) and magnesium (Z=12)) poses challenges due to the low energy of their characteristic X-rays, which are readily absorbed by air or sample matrices. To overcome this, analyses are conducted in vacuum or helium atmospheres, which minimize absorption and improve transmission efficiency—for instance, helium allows capture of over 99% of sodium's radiation compared to air.[71] These conditions extend reliable detection to elements starting from sodium, though sensitivity decreases for even lighter species like carbon or oxygen.[72]In environmental applications, XRF has been instrumental since the 1970s for assessing soil contamination by heavy metals like lead (Pb) and arsenic (As), correlating well with laboratory digestion methods for on-site screening.[73] The U.S. Environmental Protection Agency (EPA) adopted XRF for air particulate analysis in the early 1970s, developing spectrometers to quantify elemental composition on filters and monitor pollutants under emerging standards.[74] This facilitated rapid source apportionment of fine particulates, aiding compliance with Clean Air Act provisions.[75]For pharmaceutical and food safety, XRF enables non-destructive testing of heavy metals such as lead, cadmium, and mercury, ensuring products meet stringent limits without altering samples for further use.[76] In foods, it screens vegetables and other matrices for trace contaminants at ppm levels, supporting EU and U.S. regulations on elemental impurities.[77] These applications highlight XRF's role in preventive quality assurance, from raw materials to finished goods.The technique's commercial viability began in the 1950s with the introduction of the first production X-ray spectrometers, marking a milestone in accessible elemental analysis for industrial and research use.[3]
Structural and Material Characterization
X-ray fluorescence (XRF) enables phase identification in materials by analyzing ratios of characteristic X-ray lines, such as Kα to Kβ or L lines, which vary with the chemical environment and oxidation state of elements. For instance, the Kβ/Kα intensity ratio in iron-bearing minerals shifts due to differences in the 3p-3d electron interactions, allowing distinction between Fe²⁺ and Fe³⁺ species; in aqueous solutions, Fe³⁺ exhibits a distinct Kβ spectrum not reproducible by simple scaling of Fe²⁺ signals. This approach has been applied to igneous rocks, where the FeO/Fe₂O₃ total ratio is determined from K- and L-series line intensities, providing insights into mineral oxidation states without destructive sample preparation. Similarly, L-line ratios in heavier elements like uranium or platinum compounds reflect coordination chemistry, aiding compound-specific identification in complex matrices.In polycrystalline materials, XRF reveals texture through variations in fluorescence intensity arising from preferred crystallite orientations, as the excitation and emission geometry influences detected signals. When samples are rotated or tilted during analysis, anisotropic intensity patterns emerge, indicating alignment along specific lattice planes; this is particularly useful for metals and ceramics where random orientation assumptions fail. Such effects must be corrected in quantitative models, similar to those in related X-ray techniques, to avoid errors in phase abundance estimates.For thin films and coatings, XRF facilitates depth profiling and thickness measurements using grazing incidence configurations, where the X-ray beam angle is kept below the critical angle for total external reflection, enhancing surface sensitivity. This method probes depths up to several micrometers, enabling non-destructive assessment of multilayer structures; for example, gallium depth profiles in Cu(In₁₋ₓGaₓ)Se₂ thin films are quantified by modeling fluorescence yields as a function of incidence angle. Thicknesses of ultra-thin iron films, down to nanometers, have been determined alongside density by analyzing normalized K-line intensities under grazing conditions, with accuracy improved by accounting for self-absorption.In geological contexts, XRF supports provenance studies of sediments through patterns of rare earth elements (REEs), whose normalized abundances serve as fingerprints of source rocks. Portable XRF scanners detect REEs like lanthanum and cerium in sediment cores, revealing transport histories and depositional environments; for instance, variations in light versus heavy REE ratios trace anthropogenic inputs or volcanic origins in coastal deposits. This elemental proxy approach complements isotopic methods, offering rapid, in-field resolution for paleoenvironmental reconstructions.XRF integrates effectively with X-ray diffraction (XRD) for comprehensive material characterization, where XRF provides elemental composition and line-ratio-based chemical insights, while XRD identifies crystalline phases and structures. Combined systems enable simultaneous analysis of the same sample spot, enhancing accuracy in cultural heritage artifacts and mining ores by correlating elemental distributions with mineralogy. This synergy is vital for complex samples like soils, where XRF detects REE provenance indicators and XRD confirms phase assemblages.
Related Spectroscopic Techniques
Particle-induced X-ray emission (PIXE) is an analytical technique that employs accelerated charged particles, typically protons with energies in the mega-electron-volt (MeV) range, to bombard a sample and induce characteristic X-ray emission for elemental analysis. Unlike conventional X-ray fluorescence (XRF), which uses photons for excitation, PIXE leverages the higher ionization cross-sections of ions, resulting in enhanced sensitivity for trace elements, often reaching parts-per-billion (ppb) detection limits in thin samples where self-absorption effects are minimized. This makes PIXE particularly advantageous for analyzing aerosol filters, biological tissues, and cultural artifacts, with the added benefit of simultaneous multi-element detection due to the broad excitation spectrum. The technique's synergy with XRF lies in its shared reliance on characteristic X-ray spectra for identification, though PIXE's ion-induced background is lower for mid-Z elements, enabling complementary use in environmental monitoring.[78][79]Electron probe microanalysis (EPMA) integrates wavelength-dispersive X-ray fluorescence (WDXRF) spectroscopy with scanning electron microscopy (SEM) to provide high-resolution chemical mapping at the micron scale. In EPMA, a focused electron beam excites the sample, generating characteristic X-rays that are dispersed by a crystal spectrometer for precise elemental quantification, achieving spatial resolutions typically between 0.5 and 5 μm depending on beam conditions and sample conductivity. This method excels in spatially resolved analysis of heterogeneous materials, such as minerals and alloys, where it reveals elemental distributions that inform phase identification and diffusion processes. Compared to standard XRF, EPMA offers superior lateral resolution through electron beam focusing but is limited to conductive or coated samples to avoid charging effects; its synergy with XRF enables validation of bulk compositions against microscale variations in metallurgical studies.[80][81]Total reflection X-ray fluorescence (TXRF) enhances the sensitivity of XRF for trace element detection by exploiting the total external reflection of X-rays at a grazing incidence angle below the critical angle of the substrate, typically around 0.1–0.5° for common reflectors like quartz or silicon. This configuration confines the excitation and detection volume to a shallow depth of a few nanometers, drastically reducing background fluorescence from the substrate and enabling detection limits in the nanogram-per-milliliter range for liquid samples, such as environmental waters or biological fluids. TXRF is ideal for ultra-trace analysis without extensive sample preparation, as microliter volumes can be directly deposited on the reflector; its primary difference from conventional XRF is the improved signal-to-noise ratio for surface and thin-film analyses, fostering synergies in quality control for pharmaceuticals and semiconductors.[82][83]Synchrotron-based XRF, often implemented as micro-XRF (μXRF), utilizes the intense, tunable beams from synchrotron sources to achieve high-flux excitation for elemental imaging with sub-micrometer spatial resolution, frequently below 1 μm through focused optics like Kirkpatrick-Baez mirrors. This approach enables two- and three-dimensional mapping of trace elements in complex matrices, such as biological tissues or geological samples, with sensitivities down to femtograms per pixel, far surpassing laboratory XRF due to the synchrotron's brightness and lack of self-absorption in raster-scanned modes. The technique's key synergy with standard XRF is in extending bulk analysis to spatially resolved studies, such as speciation in environmental remediation, while its tunability allows selective excitation of specific edges for reduced spectral overlap.[84][85]X-ray absorption spectroscopy (XAS) complements emission-based techniques like XRF by probing the local atomic structure and electronic environment around a specific absorbing atom through measurement of X-ray transmission or fluorescence as a function of energy near absorption edges. The extended X-ray absorption fine structure (EXAFS) region of XAS provides quantitative information on coordination numbers, bond distances, and disorder in the first few coordination shells, typically up to 5–6 Å, by analyzing the interference between outgoing and backscattered photoelectrons. While XRF identifies elemental presence via emission lines, XAS elucidates chemical speciation and bonding, making it synergistic for comprehensive material characterization in catalysis and geochemistry; for instance, combining XRF mapping with XAS reveals both distribution and local geometry of metal centers in heterogeneous catalysts.[86][87]
Instrumentation and Quality Control
Sample Preparation Techniques
Sample preparation is crucial in X-ray fluorescence (XRF) analysis to ensure homogeneity, minimize matrix effects, and achieve accurate elemental quantification, as heterogeneous samples can lead to biased intensity measurements.[88] Proper techniques address challenges such as particle size variations and contamination, which can introduce errors exceeding 10-20% in trace element detection without correction.[89]For solid samples, common methods include pressing powders into pellets and preparing fusion beads. Pressed pellets are formed by grinding the sample to a fine powder (typically <75 μm) and compacting it under high pressure, often with a binder like cellulose or wax, into disks of 13 mm diameter to create a uniform surface for XRF excitation.[90] This technique is widely used for geological and cement samples, reducing heterogeneity and enabling reproducible analysis.[91] Fusion beads involve mixing the powdered sample with a flux such as lithium tetraborate (Li₂B₄O₇) or a 50:50 lithium tetraborate-lithium metaborate blend, then heating to 1000-1200°C in a platinum crucible to form a homogeneous glass disk upon cooling. This method eliminates particle size and mineralogical effects, providing superior accuracy for major and minor elements in ores and refractories.[92]Liquid samples can be analyzed directly in specialized cells with X-ray transparent windows, such as polypropylene cups, for bulk liquids like oils or solutions, though this may require dilution to avoid absorption losses.[93] Alternatively, liquids are dried on thin films (e.g., Mylar or quartz) to form solid residues, concentrating analytes for better sensitivity in environmental or biological matrices.[94] For ultra-trace analysis, total reflection X-ray fluorescence (TXRF) employs microliter volumes pipetted onto reflective carriers like silicon wafers, enabling detection limits down to ppb levels without digestion.[95]Loose powders are prepared by loading the ground material into open-ended cells sealed with thin Mylar films (typically 4-6 μm thick) to contain the sample while allowing X-ray transmission and minimizing heterogeneity during measurement.[96] This approach is suitable for rapid screening of powders like soils or pharmaceuticals but requires careful leveling to avoid uneven packing, which can cause up to 5-10% intensity variations.[97]Surface preparation for solid or coated samples involves grinding with abrasives (e.g., silicon carbide) to <50 μm particle size and polishing to a flat finish, as larger grains (>50 μm) induce secondary fluorescence and absorption errors, potentially inflating concentrations by 15% or more.[89] These steps ensure representative sampling depths of 10-100 μm, critical for alloys and thin films.[98]Certified reference materials, such as NIST Standard Reference Materials (SRMs) like SRM 2709a for soils or SRM 80a for glass, are prepared using similar techniques and used to validate sample handling, with protocols emphasizing clean tools and storage to prevent contamination from trace metals.[99] Matrix corrections may be applied post-preparation to account for inter-element effects, but their efficacy depends on initial sample uniformity.[88]
Instrument Calibration and Qualification
Instrument calibration in X-ray fluorescence (XRF) spectrometry is essential to ensure accurate and reliable quantitative analysis by establishing the relationship between measured X-ray intensities and element concentrations. Calibration procedures typically involve the use of reference materials to account for instrumental factors such as excitation efficiency, detection sensitivity, and matrix effects. Two primary methods are employed: the influence coefficients approach, particularly the α method, and the fundamental parameters (FP) method, which is widely used for multi-element analysis. The α method, introduced by Lachance and Traill in 1966, uses theoretical coefficients to correct for absorption effects in binary systems, expressed as C_i = R_i (1 + \sum_j \alpha_{ij} C_j), where C_i is the concentration of element i, R_i is the relative intensity, and \alpha_{ij} represents the influence of element j on i, assuming monochromatic excitation and ignoring enhancement.[100] This empirical-theoretical hybrid is suitable for simpler matrices but has limitations in accuracy for complex samples due to its neglect of secondary fluorescence.[101] In contrast, the FP method relies on physical models derived from Sherman's equation to predict intensities based on atomic parameters like fluorescence yield and mass absorption coefficients, enabling robust multi-element quantification without extensive empirical standards, though it requires accurate knowledge of instrumental geometry and spectral distribution.[102]Traceability of XRF measurements to international standards is mandated by ISO/IEC 17025:2017, which requires laboratories to demonstrate metrological traceability through a documented unbroken chain of calibrations or comparisons to certified reference materials (CRMs) with stated uncertainties.[103] CRMs, such as those provided by NIST (e.g., Standard Reference Materials for alloys and geological samples), are used to validate calibration curves and ensure comparability across instruments, with values certified via independent techniques like inductively coupled plasma mass spectrometry.[104] For instance, NIST SRM 2709 (San Joaquin Soil) serves as a traceable standard for soil and sedimentanalysis in XRF.[105]Instrument qualification involves periodic tests to verify performance parameters, including energy resolution and long-term stability. Energy resolution, defined as the full width at half maximum (FWHM) of a characteristic X-ray peak, is assessed by measuring a monochromatic line (e.g., Mn Kα at 5.9 keV) to ensure the detector can distinguish overlapping peaks, typically targeting <150 eV for silicon drift detectors in energy-dispersive XRF (EDXRF).[15] Stability is evaluated through drift monitors, such as stable glass discs, to check intensity variations; modern XRF spectrometers exhibit drift less than 0.1% over 24 hours due to high reproducibility of X-ray tubes and detectors.[106] These tests confirm operational integrity before analysis.Uncertainty estimation in XRF quantifies contributions from random and systematic sources to provide confidence intervals for results. Counting statistics follow Poisson distribution, yielding a relative uncertainty of $1 / \sqrt{N}, where N is the number of counts, dominating at low concentrations.[41] Matrix effects, including absorption and enhancement, are incorporated via weighted least squares regression, using Compton scatter intensity as an internal standard to predict overall uncertainty as u(x_{pred}) = \sqrt{s_{samp}^2 + s_{calib}^2}, where terms account for sample variance and calibration errors relative to a reference matrix.[107]Software validation for quantitative XRF analysis follows ASTM E1621, a standard guide for wavelength-dispersive XRF that outlines procedures for method development, including calibrationverification and precision checks to achieve acceptable accuracy (typically <1-2% relative error for major elements).[108] This involves testing against CRMs to confirm the software's handling of matrix corrections and intensity calculations, ensuring compliance with ISO 17025 for accredited laboratories.[109]
Modern Developments and Advances
Since the early 2000s, portable and handheld X-ray fluorescence (XRF) spectrometers have revolutionized on-site elemental analysis by incorporating battery-powered, compact designs that enable non-destructive testing in remote environments.[110] These devices, often weighing less than 2 kg, utilize low-power X-ray tubes (typically 5-12 W) and silicon drift detectors for real-time data acquisition, with integrated software applications facilitating immediate interpretation during field archaeology surveys, such as sourcing obsidian artifacts in the Mediterranean.[111][112]Benchtop micro-XRF systems have advanced spatial resolution to spot sizes below 50 μm, enabling high-resolution 2D elemental mapping of heterogeneous samples without sample preparation.[113] For instance, the Orbis series achieves 30-60 μm spots using poly-capillary optics, allowing detailed analysis of layered structures in art conservation, such as mapping silver distribution in historic photographs to assess degradation.[114] Similarly, the ATLAS Apex M provides down to 5 μm resolution for hyperspectral imaging in material characterization.[115]The integration of artificial intelligence and machine learning has significantly enhanced XRF spectrum processing, particularly for deconvoluting overlapping peaks that challenge traditional methods. Deep learning algorithms, trained on synthetic datasets generated via Monte Carlo simulations, accurately separate closely spaced lines like Fe-Kα and Mn-Kβ (at 6.40 and 6.49 keV), reducing artifacts in elemental maps and improving quantification in low-signal regions.[116] The AGA-BP-Attention neural network, optimized by adaptive genetic algorithms, further addresses non-linear backgrounds and peak overlaps in complex spectra, achieving 95.75% recognition accuracy for multi-elemental samples and suppressing noise to enhance limits of detection to sub-ppm levels.[117]Efforts toward sustainability in XRF instrumentation include the development of low-power excitation sources and eco-friendly anode materials to reduce energy consumption and environmental impact. Air-cooled, metal-ceramic X-ray tubes operating at reduced voltages (e.g., below 50 kV) enable portable systems with minimal power draw while maintaining analytical performance, supporting greener field operations.[118]Anode materials like rhodium and silver, chosen for their efficiency in light element excitation, offer longer tube lifespans (up to 10,000 hours) compared to traditional tungsten, minimizing material waste and replacement frequency in sustainable designs.[119]In the 2020s, XRF has seen integration with hyperspectral imaging for multimodal analysis, combining elemental data with molecular spectroscopy to produce comprehensive 3D chemical maps of samples.[120]