(Translated by https://www.hiragana.jp/)
Element-specific, non-destructive profiling of layered heterostructures

Element-specific, non-destructive profiling of layered heterostructures

Nicolò D’Anna ndanna@ucsd.edu Present address: Department of Physics University of California San Diego, La Jolla, CA 92093, USA PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland Laboratory for Solid State Physics and Quantum Center, ETH Zurich, Zurich, Switzerland    Jamie Bragg London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Electronic and Electrical Engineering, University College London, London WC1E 7SE, UK    Elizabeth Skoropata    Nazareth Ortiz Hernández PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland    Aidan G. McConnell    Maël Clémence PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland Laboratory for Solid State Physics and Quantum Center, ETH Zurich, Zurich, Switzerland    Hiroki Ueda PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland    Procopios C. Constantinou PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Physics and Astronomy, University College London, WC1E 6BT, London, UK    Kieran Spruce London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Electronic and Electrical Engineering, University College London, London WC1E 7SE, UK    Taylor J.Z. Stock London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Electronic and Electrical Engineering, University College London, London WC1E 7SE, UK    Sarah Fearn London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Materials, Imperial College of London, London SW7 2AZ, UK    Steven R. Schofield London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Physics and Astronomy, University College London, WC1E 6BT, London, UK    Neil J. Curson London Centre for Nanotechnology, University College London, WC1H 0AH, London, UK Department of Electronic and Electrical Engineering, University College London, London WC1E 7SE, UK    Dario Ferreira Sanchez    Daniel Grolimund    Urs Staub    Guy Matmon    Simon Gerber PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland    Gabriel Aeppli gabriel.aeppli@psi.ch PSI Center for Photon Science, Paul Scherrer Institute, 5232 Villigen PSI, Switzerland Laboratory for Solid State Physics and Quantum Center, ETH Zurich, Zurich, Switzerland Institute of Physics, EPF Lausanne, 1015 Lausanne, Switzerland
Abstract

Fabrication of semiconductor heterostructures is now so precise that metrology has become a key challenge for progress in science and applications. It is now relatively straightforward to characterize classic III-V and group IV heterostructures consisting of slabs of different semiconductor alloys with thicknesses of similar-to\sim5 nm and greater using sophisticated tools such as X-ray diffraction, high energy X-ray photoemission spectroscopy, and secondary ion mass spectrometry. However, profiling thin layers with nm or sub-nm thickness, e.g. atomically thin dopant layers (δでるた𝛿\deltaitalic_δでるた-layers), of impurities required for modulation doping and spin-based quantum and classical information technologies is more challenging. Here, we present theory and experiment showing how resonant-contrast X-ray reflectometry meets this challenge. The technique takes advantage of the change in the scattering factor of atoms as their core level resonances are scanned by varying the X-ray energy. We demonstrate the capability of the resulting element-selective, non-destructive profilometry for single arsenic δでるた𝛿\deltaitalic_δでるた-layers within silicon, and show that the sub-nm electronic thickness of the δでるた𝛿\deltaitalic_δでるた-layers corresponds to sub-nm chemical thickness. In combination with X-ray fluorescence imaging, this enables non-destructive three-dimensional characterization of nano-structured quantum devices. Due to the strong resonances at soft X-ray wavelengths, the technique is also ideally suited to characterize layered quantum materials, such as cuprates or the topical infinite-layer nickelates.

With the advent of scanning tunneling microscopy (STM) lithography in the 1990s [1, 2], it became possible to fabricate dopant-based nano-electronic structures in semiconductors [3], as can now be done with phosphorus, arsenic, and boron in silicon [4, 5, 6]. In the meantime, industrially fabricated transistors have reached a 7 nm scale [7]. A number of methods to image electronic nano-structures are available. The most popular are destructive, and include transmission electron microscopy [8], atom-probe tomography [9, 8] and secondary ion mass spectrometry (SIMS) [10, 8]. Non-destructive imaging techniques include X-ray fluorescence [11, 12], X-ray diffraction [13], angle-resolved photoemission spectroscopy (ARPES) [14], ellipsometry [15, 16], as well as scanning microwave [17, 18], broadband electrostatic force [19], and single-electron probe [20] microscopy. Additionally, imaging techniques based on ion-beams, such as nuclear reaction analysis [21], medium energy ion and Rutherford scattering [22, 23, 24], can be non-destructive for conductive layers but are typically slow. Most of these non-destructive techniques only give two-dimensional lateral information, whilst X-ray methods can produce three-dimensional images by tomography [25, 26] at the expense of time. Conversely, provided strong refractive index contrast, X-ray reflectometry [27] can measure the vertical depth-profile of atomically thin dopant layers in a reasonably short time, of order 10 minutes per scan already since the 1990s [28]. But for thin layers with small dopant concentrations such as atomically thin layers of dopant atoms (δでるた𝛿\deltaitalic_δでるた-layers), considerable modelling is required to determine the dimensions of the layers, particularly in the presence of features such as surface roughness and oxidation or other elements in the device grown on a substrate wafer.

X-ray reflectometry can be made more sensitive to a specific element by measuring resonantly at energies around the corresponding X-ray absorption edge [29]. The atomic resonance induces a large phase shift in the reflected signal, which can be used to isolate the element’s contribution to the reflectivity and obtain its distribution as a function of depth. This is particularly relevant for low concentrations, as in dopant-defined devices in silicon which are based on δでるた𝛿\deltaitalic_δでるた-layers with 3D dopant concentrations typically below 5% [5]. Here we show that for a δでるた𝛿\deltaitalic_δでるた-layer of arsenic donors in silicon, analysis of resonant contrast X-ray reflectometry (RCXR) can be enormously simplified, requiring no modelling of the δでるた𝛿\deltaitalic_δでるた-layer’s host heterostructure. Furthermore, for the devices studied here SIMS gives layer thicknesses which are >2absent2>2> 2 nm [14], considerably larger than the typically <1absent1<1< 1 nm thicknesses provided by measures accessing the conduction electrons in the δでるた𝛿\deltaitalic_δでるた-layers [5, 14, 16]; in this work reflectometry resolves this discrepancy by showing that the chemical (arsenic) layer thicknesses are considerably smaller than the SIMS results and consistent with the electronic thicknesses established by ARPES and magneto-resistance. Therefore, RCXR is well suited for non-destructive high-throughput preliminary characterizations of semiconductor heterostructures.

Resonant-contrast X-ray reflectometry

Refer to caption
Figure 1: X-ray reflectometry. Schematic of the sample’s layer structure and the measurement geometry. The surface consists of oxide (SiO2) and silicon with a combined thickness of d=dSiO2+dSi𝑑subscript𝑑subscriptSiO2subscript𝑑Sid=d_{\rm SiO_{2}}+d_{\rm Si}italic_d = italic_d start_POSTSUBSCRIPT roman_SiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_d start_POSTSUBSCRIPT roman_Si end_POSTSUBSCRIPT. Within the silicon lattice there is an arsenic-doped layer of thickness δでるた𝛿\deltaitalic_δでるた. X-rays shine on the sample with an incidence angle θしーた𝜃\thetaitalic_θしーた (with respect to the surface) and the specular reflection is detected by a detector placed at an angle θしーたD=2θしーたsubscript𝜃𝐷2𝜃\theta_{D}=2\thetaitalic_θしーた start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT = 2 italic_θしーた (with respect to the incident beam). Constructive interference occurs at 2dsinθしーた=mλらむだ2𝑑𝜃𝑚𝜆2d\sin\theta=m\lambda2 italic_d roman_sin italic_θしーた = italic_m italic_λらむだ for two interfaces separated by a distance d𝑑ditalic_d, where m𝑚mitalic_m is an integer and λらむだ𝜆\lambdaitalic_λらむだ is the X-ray wavelength.

X-ray reflectometry was first used in 1954 [30] to measure the thickness of copper on glass and since then has become a common technique for studying a wide variety of layered materials, including surfaces, thin films and multilayers [31]. It involves measuring the specular reflection from sample surfaces (see Fig. 1). The angle θしーた𝜃\thetaitalic_θしーた is swept to obtain a reflectometry measurement.

Light travelling through a medium is scattered upon changes in the medium’s scattering length density ρろー=r0qNqfq𝜌subscript𝑟0subscript𝑞subscript𝑁𝑞subscript𝑓𝑞\rho=r_{0}\sum_{q}N_{q}f_{q}italic_ρろー = italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT, where r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the classical electron radius, Nqsubscript𝑁𝑞N_{q}italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT and fq=f1,q+if2,qsubscript𝑓𝑞subscript𝑓1𝑞𝑖subscript𝑓2𝑞f_{q}=f_{1,q}+if_{2,q}italic_f start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 1 , italic_q end_POSTSUBSCRIPT + italic_i italic_f start_POSTSUBSCRIPT 2 , italic_q end_POSTSUBSCRIPT are the number of atoms per unit volume and the complex atomic scattering factor for an atom of element q𝑞qitalic_q, respectively. For typical semiconductors X-rays scatter weakly, hence, for a continuously varying depth profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ), only single scattering events need to be considered and the measured reflection is described by the kinematic Born approximation [32, 33, 27],

R(Q)=(4πぱい)2Q2|ρろー(z)eiQzdz|2=rF(Q)2|FQ(ρろー)|2,𝑅𝑄superscript4𝜋2superscript𝑄2superscript𝜌𝑧superscript𝑒𝑖𝑄𝑧differential-d𝑧2subscript𝑟𝐹superscript𝑄2superscriptsubscript𝐹𝑄𝜌2R(Q)=\frac{(4\pi)^{2}}{Q^{2}}\left|\int\rho(z)e^{-iQz}\ \mathrm{d}z\right|^{2}% =r_{F}(Q)^{2}|F_{Q}(\rho)|^{2},italic_R ( italic_Q ) = divide start_ARG ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG | ∫ italic_ρろー ( italic_z ) italic_e start_POSTSUPERSCRIPT - italic_i italic_Q italic_z end_POSTSUPERSCRIPT roman_d italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (1)

where Q=4πぱいsinθしーた/λらむだ𝑄4𝜋𝜃𝜆Q=4\pi\sin\theta/\lambdaitalic_Q = 4 italic_πぱい roman_sin italic_θしーた / italic_λらむだ is the Q𝑄Qitalic_Q-vector, λらむだ𝜆\lambdaitalic_λらむだ the X-ray wavelength, FQ(x)subscript𝐹𝑄𝑥F_{Q}(x)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_x ) the Fourier transform, and rF(Q)2=(4πぱい)2Q2subscript𝑟𝐹superscript𝑄2superscript4𝜋2superscript𝑄2r_{F}(Q)^{2}=\frac{(4\pi)^{2}}{Q^{2}}italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG the squared Fresnel reflectivity.

The scattering length density profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ) of a material is the sum of each of its elements’ profiles qρろーq(z)subscript𝑞subscript𝜌𝑞𝑧\sum_{q}\rho_{q}(z)∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_ρろー start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_z ). Therefore, if an element q𝑞qitalic_q is dilute, it can be considered to contribute perturbatively to the host profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ) and, thus, to Eq. (1). So expanding R(Q)𝑅𝑄R(Q)italic_R ( italic_Q ) in terms of ρろーq(z)subscript𝜌𝑞𝑧\rho_{q}(z)italic_ρろー start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_z ), the first-order term will be due to interference between the element q𝑞qitalic_q and the host, represented mathematically by a product of the Fourier transforms of ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ) and the dilute elemental density profile denoted δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ) rather than the square of Fourier amplitude of δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ) appearing to second order. In the first part of our Methods section, we develop an analytic theoretical description with relevant equations that describe the consequences, with the following key conclusions:

  1. 1.

    Reflectometry data collected at a single X-ray energy can be separated into components belonging to different length-scales by Fourier filtering at the appropriate frequencies. As a consequence, if the dilute layer of element q𝑞qitalic_q is thin compared to the other layers, its signal can readily be isolated, and if it has a Gaussian profile (of standard deviation σしぐま𝜎\sigmaitalic_σしぐま) its contribution is given by:

    I(Q)|F|LF2=2r0N2D|Δでるたf|exp((Qσしぐま)22)cos(Qdϕ),𝐼𝑄subscriptsuperscript𝐹2LF2subscript𝑟0subscript𝑁2DΔでるた𝑓superscript𝑄𝜎22𝑄𝑑italic-ϕ\frac{I(Q)}{\sqrt{|F|^{2}_{\mathrm{LF}}}}=2r_{0}N_{\text{2D}}|\Delta f|\exp% \left(-\frac{(Q\sigma)^{2}}{2}\right)\cos(Qd-\phi),divide start_ARG italic_I ( italic_Q ) end_ARG start_ARG square-root start_ARG | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT end_ARG end_ARG = 2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT | roman_Δでるた italic_f | roman_exp ( - divide start_ARG ( italic_Q italic_σしぐま ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) roman_cos ( italic_Q italic_d - italic_ϕ ) , (2)

    where d𝑑ditalic_d and N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT are the dilute layer’s depth and 2D density, respectively, δでるた=22πぱいσしぐま𝛿22𝜋𝜎\delta=2\sqrt{2}\pi\sigmaitalic_δでるた = 2 square-root start_ARG 2 end_ARG italic_πぱい italic_σしぐま the layer’s thickness, Δでるたf=fqfhostΔでるた𝑓subscript𝑓𝑞subscript𝑓host\Delta f=f_{q}-f_{\text{host}}roman_Δでるた italic_f = italic_f start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT host end_POSTSUBSCRIPT, ϕ=arg(Δでるたf)italic-ϕargΔでるた𝑓\phi=\text{arg}(\Delta f)italic_ϕ = arg ( roman_Δでるた italic_f ), I(Q)𝐼𝑄I(Q)italic_I ( italic_Q ) the first-order term of the Born approximation’s expansion in δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ) [see Eq. (6) and (8)], and |F|LFsubscript𝐹LF|F|_{\mathrm{LF}}| italic_F | start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT the low frequency part of the reflectometry.

  2. 2.

    Taking the difference ΔでるたR(Q)Δでるた𝑅𝑄\Delta R(Q)roman_Δでるた italic_R ( italic_Q ) between two reflectometry measurements at energies E𝐸Eitalic_E and Esuperscript𝐸E^{\prime}italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT straddling a resonant edge of the dilute layer’s element species, isolates the signal from the dilute layer [see Eq. (13) and (14)]. For a Gaussian profile this leads to Eq. (2) with the substitutions I(Q)ΔでるたR(Q)/rF(Q)2𝐼𝑄Δでるた𝑅𝑄subscript𝑟𝐹superscript𝑄2I(Q)\to\Delta R(Q)/r_{F}(Q)^{2}italic_I ( italic_Q ) → roman_Δでるた italic_R ( italic_Q ) / italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and ΔでるたfΔでるたfEΔでるたfEΔでるた𝑓Δでるたsubscript𝑓superscript𝐸Δでるたsubscript𝑓𝐸\Delta f\to\Delta f_{E^{\prime}}-\Delta f_{E}roman_Δでるた italic_f → roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT - roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT, which directly yields the depth d𝑑ditalic_d and thickness δでるた𝛿\deltaitalic_δでるた of the layer, as opposed to conventional fitting of the full reflectometry data.

This first-order perturbation theory has the benefit of measuring the dimensions of a dilute layer within a heterostructure of other elements without the need to model the full structure. An ideal test-bed for this are the substitutional, nm-thin dopant δでるた𝛿\deltaitalic_δでるた-layer in silicon, with typical 2D density N2Dsubscript𝑁2DabsentN_{\text{2D}}\approxitalic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT ≈1014 cm-2 corresponding to N3D=N2D/δでるたsubscript𝑁3Dsubscript𝑁2D𝛿absentN_{\text{3D}}=N_{\text{2D}}/\delta\approxitalic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT / italic_δでるた ≈1017 cm-3 (for δでるたsimilar-to𝛿absent\delta\simitalic_δでるた ∼1 nm) representing <<<5% of the total silicon atoms [5].

Refer to caption (a)

Refer to caption (b)

Figure 2: Dopant contribution to reflectometry across the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge. (a) X-ray reflectometry of sample #1 with photon energies below (red, 1300 eV) and above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-absorption edge (black, 1335 eV), obtained by rotating the sample and the detector to measure the specular reflection. The horizontal axis shows the wave-vector change on scattering Q𝑄Qitalic_Q, to directly compare the interference patterns of both energies. The inset depicts the fast Fourier transform (FFT) of the reflectometry curves on a logarithmic scale, directly yielding the depth d𝑑ditalic_d of the arsenic δでるた𝛿\deltaitalic_δでるた-layer. The right axis of the inset shows the FFT phase and the phase difference (blue) at both energies. (b) The black line shows the R𝑅Ritalic_R(1335 eV) data from (a) multiplied by Q2superscript𝑄2Q^{2}italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. A Fourier transform filter was used to separate frequencies corresponding to reflections from depths less than 10 nm (R𝑅Ritalic_R(Si), green) and greater than 10 nm (R𝑅Ritalic_R(As), blue). The gray line represents the difference ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R between the reflectometry measured above (1335 eV) and below (1300 eV) the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge. Both R𝑅Ritalic_R(As) and ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R have been divided by R(Si)𝑅Si\sqrt{R(\text{Si})}square-root start_ARG italic_R ( Si ) end_ARG to isolate the arsenic contribution. The inset shows the arsenic scattering length density ρろーAssubscript𝜌As\rho_{\text{As}}italic_ρろー start_POSTSUBSCRIPT As end_POSTSUBSCRIPT, obtained from R𝑅Ritalic_R(As) and ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R as described by Eq. (11) in the Methods. Dotted curves denote fits using the DYNA program and Eq. (2) in (a) and (b), respectively.

Extracting the arsenic signal from reflectometry

The samples used to demonstrate our method are Si(001) wafers with arsenic δでるた𝛿\deltaitalic_δでるた-layers (Si:As) at varying depths d=1575𝑑1575d=15-75italic_d = 15 - 75 nm below the surface and an oxide (SiO2) surface layer of width dSiO21subscript𝑑subscriptSiO21d_{\rm SiO_{2}}\approx 1italic_d start_POSTSUBSCRIPT roman_SiO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≈ 1 nm [34] (see Fig. 1). They were prepared as described in the Methods, with resulting two-dimensional dopant density N2D1014subscript𝑁2Dsuperscript1014N_{\mathrm{2D}}\approx 10^{14}italic_N start_POSTSUBSCRIPT 2 roman_D end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT cm-2, as determined by SIMS, STM, magneto-resistance, and X-ray fluorescence [5, 11]. In addition, one control sample features a buried oxide layer instead of arsenic.

In the doped region, considering a dAs1subscript𝑑As1d_{\rm As}\approx 1italic_d start_POSTSUBSCRIPT roman_As end_POSTSUBSCRIPT ≈ 1 nm thick layer, <5%absentpercent5<5\%< 5 % of silicon atoms are replaced by arsenic, while the oxide contains twice as many oxygen atoms as silicon. Therefore, the optical contrast between the silicon and Si:As is small compared to that between oxide and silicon. Figure 2a shows typical reflectometry data for our samples. The black data are for an energy of 1335 eV, above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-absorption edge at 1324 eV [35], and the red ones for 1300 eV, below the edge. Both datasets contain distinct fast oscillations due to interference between reflections from the δでるた𝛿\deltaitalic_δでるた-layer and those from the surface region. The periodicity of the oscillations is related to the dopant layer depth d𝑑ditalic_d through the Bragg condition d=mλらむだ/2sinθしーた𝑑𝑚𝜆2𝜃d=m\lambda/2\sin\thetaitalic_d = italic_m italic_λらむだ / 2 roman_sin italic_θしーた (corrections from the refractive index n𝑛nitalic_n can be omitted for X-rays), i.e., a smaller d𝑑ditalic_d implies a larger period. In the inset of Fig. 2a, a depth of d=18𝑑18d=18italic_d = 18 nm is obtained from the Fourier transform. The fast oscillations are modulated by an envelope that varies slowly with Q𝑄Qitalic_Q on a scale inversely proportional to the thin arsenic layer thickness δでるた𝛿\deltaitalic_δでるた and broadened also by the surface roughness.

At very small angles (i.e. for Q<1.5𝑄1.5Q<1.5italic_Q < 1.5 nm-1) the signal contains no meaningful structure on account of a combination of total external reflection and the leakage of light directly to the detector due to growth of the beam footprint to larger than the sample. At higher angles we can analyze the data quantitatively: the oscillations decay because the overall signal intensity decreases by the factors rF(Q)2Q2proportional-tosubscript𝑟𝐹superscript𝑄2superscript𝑄2r_{F}(Q)^{2}\propto Q^{-2}italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∝ italic_Q start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and H(Q)𝐻𝑄H(Q)italic_H ( italic_Q ) [see Eq. (9) in Methods]. Surface roughness reduces the intensity by an additional factor which can be modelled by eQ2σしぐまr2superscript𝑒superscript𝑄2superscriptsubscript𝜎𝑟2e^{-Q^{2}\sigma_{r}^{2}}italic_e start_POSTSUPERSCRIPT - italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_σしぐま start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, where σしぐまrsubscript𝜎𝑟\sigma_{r}italic_σしぐま start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the root mean square roughness [36, 33]. This roughness was measured with STM and atomic force microscopy (AFM, see Methods), and found to be 0.1 nm, consistent with other studies [37].

Refer to caption (a)

Refer to caption (b)

Figure 3: Resonant reflectometry across the arsenic L2,3subscript𝐿23L_{2,3}italic_L start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT-edges. (a) Sharp reflectometry edge at the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-absorption edge resonance of 1324 eV in sample #1, at an incidence angle of θしーた=10𝜃superscript10\theta=10^{\circ}italic_θしーた = 10 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, normalised by the incident photon flux I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The blue line is a fit assuming δでるた𝛿\deltaitalic_δでるた = 1.5 nm (see Methods and Fig. 5). The top inset depicts the resonance process where an X-ray photon is absorbed and excites an electron from the L𝐿Litalic_L-shell to the unoccupied M𝑀Mitalic_M- or N𝑁Nitalic_N-shells. The bottom inset depicts the energy dependence of the arsenic, silicon, and oxygen scattering factors in full, dashed, and dotted lines, respectively [38]. Only the arsenic scattering factor changes noticeably in this energy range. (b) Simulated (black) and experimental (red) relative reflection change ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R, assuming N2D=2.77×1014subscript𝑁2D2.77superscript1014N_{\text{2D}}=2.77\times 10^{14}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 2.77 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT cm-2. The black line is a guide to the eye used to estimate the error of the arsenic layer thickness δでるた𝛿\deltaitalic_δでるた (see Methods). The intersection of the data and the simulation gives δでるた=1.6±0.5𝛿plus-or-minus1.60.5\delta=1.6\pm 0.5italic_δでるた = 1.6 ± 0.5 nm for sample #1.

The phase between the two signals, shown in the inset of Fig. 2a is shifted at the arsenic resonance edge, as described by Eq. (6), where the second term is phase sensitive and proportional to the arsenic scattering length density. The expected phase shift is given by the scattering factor change Δでるたf=fAsfSiΔでるた𝑓subscript𝑓Assubscript𝑓Si\Delta f=f_{\text{As}}-f_{\text{Si}}roman_Δでるた italic_f = italic_f start_POSTSUBSCRIPT As end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT Si end_POSTSUBSCRIPT, as shown in equation Eq. (8). Thus, reflectometry data taken at 1300 eV and 1335 eV will be shifted by arg[Δでるたf(1335eV)]arg[Δでるたf(1300eV)](0.33±0.10)πぱいargdelimited-[]Δでるた𝑓1335eVargdelimited-[]Δでるた𝑓1300eVplus-or-minus0.330.10𝜋\text{arg}\left[\Delta f(1335\text{eV})\right]-\text{arg}\left[\Delta f(1300% \text{eV})\right]\approx(0.33\pm 0.10)\piarg [ roman_Δでるた italic_f ( 1335 eV ) ] - arg [ roman_Δでるた italic_f ( 1300 eV ) ] ≈ ( 0.33 ± 0.10 ) italic_πぱい, using scattering factors from [38]. This is in agreement with the measured phase shift of (0.27±0.03)πぱいplus-or-minus0.270.03𝜋(0.27\pm 0.03)\pi( 0.27 ± 0.03 ) italic_πぱい at 18 nm indicated by the peak in the inset of Fig. 2a.

Conventional fits to the data using the DYNA [39] software in Fig. 2a are shown with dotted lines, where the three-dimensional density N3Dsubscript𝑁3DN_{\text{3D}}italic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT and the δでるた𝛿\deltaitalic_δでるた-layer thickness are obtained from fitting the resonance in Fig. 3, to reduce the number of free fitting parameters. Direct fits to the data for a single photon energy are unable to give reliable results for the δでるた𝛿\deltaitalic_δでるた-layer thickness δでるた𝛿\deltaitalic_δでるた, due to the large fitting parameter space containing each layer’s thickness, density, and roughness metrics, highlighting the need for a dedicated method to extract δでるた𝛿\deltaitalic_δでるた.

To analyze the data at an X-ray energy of 1335 eV, above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge (black data in Fig. 2a), the data are multiplied by Q2/(4πぱい)2superscript𝑄2superscript4𝜋2Q^{2}/(4\pi)^{2}italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to remove the Q2superscript𝑄2Q^{-2}italic_Q start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT divergence expected from Eq. (1). Thereafter, a Fourier transform filter is used to separate oscillations corresponding to reflections from depths less than 10 nm that we attribute to the silicon layer (R𝑅Ritalic_R(Si), shown in green in Fig. 2b), and oscillations that stem from depths greater than 10 nm, which after division by R(Si)𝑅Si\sqrt{R(\text{Si})}square-root start_ARG italic_R ( Si ) end_ARG we attribute to the arsenic layer (R𝑅Ritalic_R(As), shown in blue in Fig. 2b). Under the assumption that the arsenic layer has a Gaussian profile, R𝑅Ritalic_R(As) can be fitted to Eq. (2) (dotted blue line, Fig. 2b) to obtain the arsenic layer depth and thickness, here d=18.1±0.1𝑑plus-or-minus18.10.1d=18.1\pm 0.1italic_d = 18.1 ± 0.1 nm and δでるた=0.9±0.2𝛿plus-or-minus0.90.2\delta=0.9\pm 0.2italic_δでるた = 0.9 ± 0.2 nm, respectively.

In the Methods section, we show that for a thin arsenic layer, the slow Q𝑄Qitalic_Q oscillations (see Fig. 2b, green) are largely due to the amplitude |FQ(ρろー)|subscript𝐹𝑄𝜌|F_{Q}(\rho)|| italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | for the unperturbed host. Thus, we obtain the inverse Fourier transform of the arsenic factor FQ(δでるたρろー(z))subscript𝐹𝑄𝛿𝜌𝑧F_{Q}(\delta{\rho}(z))italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_δでるた italic_ρろー ( italic_z ) ) by first subtracting |FQ(ρろー)|subscript𝐹𝑄𝜌|F_{Q}(\rho)|| italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | from the raw data divided by Q2/(4πぱい)2superscript𝑄2superscript4𝜋2Q^{2}/(4\pi)^{2}italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and then dividing by |FQ(ρろー)|subscript𝐹𝑄𝜌\sqrt{|F_{Q}(\rho)|}square-root start_ARG | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | end_ARG (see Fig. 2b, blue). Finally, taking the Fourier transform of |FQ(δでるたρろー(z))|subscript𝐹𝑄𝛿𝜌𝑧|F_{Q}(\delta{\rho}(z))|| italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_δでるた italic_ρろー ( italic_z ) ) | results in the scattering length density profile ρろーAs(z)subscript𝜌As𝑧\rho_{\text{As}}(z)italic_ρろー start_POSTSUBSCRIPT As end_POSTSUBSCRIPT ( italic_z ) [Eq. (11)] shown in the inset of Fig. 2b, with a large density at the expected arsenic layer depth. The resolution in z𝑧zitalic_z is similar-to\sim1.5 nm and dictated by the sampling range in Q𝑄Qitalic_Q; too low to resolve the shape of a δでるた<1𝛿1\delta<1italic_δでるた < 1 nm dopant-layer profile.

RCXR isolation of the δでるた𝛿\deltaitalic_δでるた-layer signal

To quantify the modulation of the fast oscillations in the reflectivity, we look at the difference measured below and above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge resonance energy at 1300 eV and 1335 eV, respectively, similarly to how magnetism can be extracted from the difference between left- and right-handed circularly polarized X-rays [40, 41].

Figure 2b shows in gray the difference ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R of the reflectivities at the two energies divided by square root of the silicon contribution. Since the quantity of interest is the Fourier transform of the density profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ), the data were multiplied by Q2/(4πぱい)2superscript𝑄2superscript4𝜋2Q^{2}/(4\pi)^{2}italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT consistent with Eq. (6). The subtraction removes the zeroth-order term due to the oxide and silicon layers and keeps the first-order term due to the arsenic-doped layer, because the scattering factors of the former are almost constant in this energy range, while they change drastically for the latter (see inset of Fig. 3b). The division by the Si/SiO2 contribution leaves a fast oscillation, with a period corresponding to the arsenic δでるた𝛿\deltaitalic_δでるた-layer depth, modulated by a slow envelope with a decay corresponding to the arsenic δでるた𝛿\deltaitalic_δでるた-layer thickness (described by Eq. (13)). The same slow modulation of the fast oscillation is present in the data for R(As)𝑅AsR(\text{As})italic_R ( As ) in Fig. 2b. Assuming a Gaussian arsenic density profile, the data are fitted to Eq. (2) yielding a thickness of δでるた=0.60.2+0.7𝛿superscriptsubscript0.60.20.7\delta=0.6_{-0.2}^{+0.7}italic_δでるた = 0.6 start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.7 end_POSTSUPERSCRIPT (0.41.30.41.30.4-1.30.4 - 1.3) nm and depth of d=18.4±0.1𝑑plus-or-minus18.40.1d=18.4\pm 0.1italic_d = 18.4 ± 0.1 nm in agreement with the result from a single energy (see Fig. 2b, gray), where δでるた𝛿\deltaitalic_δでるた is taken to be the full width at half maximum of the Gaussian profile and the uncertainty indicates the fit’s 95% confidence interval. We emphasise that the value extracted here represents an upper bound on the arsenic layer thickness due to the high-angle cutoff due to noise and interface roughness [35, 42, 43]. Surface roughness contributions add in quadrature to the result, here roughness was found to be of the order of 0.1 nm by STM and AFM measurements (see Methods) and is, therefore, negligible.

The Fourier transform of the ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R data in Fig. 2b also leads to the arsenic profile ρろーAs(z)subscript𝜌As𝑧\rho_{\text{As}}(z)italic_ρろー start_POSTSUBSCRIPT As end_POSTSUBSCRIPT ( italic_z ) (see inset of Fig. 2b). The obtained profile is in good agreement with that obtained from a single scan of the reflectivity, suggesting that our Fourier filtering procedures for single X-ray energies, not required when we are taking differences between data collected at different X-ray energies, are adequate. Furthermore, the resonance subtraction method suppresses fluctuations on the sides of the δでるた𝛿\deltaitalic_δでるた-layer peak.

RCXR experiments were performed on five different samples, of which four contained Si:As δでるた𝛿\deltaitalic_δでるた-layers and one reference sample a buried oxide layer. They were also measured with SIMS (see Methods). The results for both techniques are shown in Fig. 4, where the data in red are obtained from Fourier filtered data at a single energy above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge and the data in black from the resonant contrast method.

Our isolation of the δでるた𝛿\deltaitalic_δでるた-layer signal based on fitting the amplitude modulation of the dopant layer has the advantage that it relies neither on extensive multi-variable fitting, nor on prior knowledge or hypotheses concerning the sample composition other than assuming Gaussian profiles. Additionally, for thin layers, the amplitude modulation occurs over a large Q𝑄Qitalic_Q-range, such that the estimation of the layer thickness has high accuracy.

Refer to caption (a)

Refer to caption (b)

Figure 4: Comparison of dopant layer thicknesses and depths. The thickness δでるた𝛿\deltaitalic_δでるた (a) and depth d𝑑ditalic_d (b) of the Si:As layer below the surface, obtained by RCXR (black) and a single energy above the L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge (E+, red), as well as SIMS are plotted on the vertical and horizontal axes, respectively. The inset in panel (b) shows the dependence of δでるた𝛿\deltaitalic_δでるた on d𝑑ditalic_d as measured with X-ray reflectometry (XRR). The shaded regions in (a) denote the thickness range obtained by ARPES and magneto-resistance (MR) of samples prepared in the same way. SIMS depth errors are from the uncertainty in the sputter rate (see Methods). XRR errors for E+ and RCXR are taken from the 95% confidence interval of fits to Eq. (2).

Arsenic L2,3subscript𝐿23L_{2,3}italic_L start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT-edge resonance measurements

The previous section shows that upper bounds on δでるた𝛿\deltaitalic_δでるた-layer thicknesses are readily obtained from examination of the rapid oscillations associated with their depth. On the other hand, short wavelength disorder can introduce noise at high momentum transfers Q𝑄Qitalic_Q and, thereby, complicate the determination of lower bounds of δでるた𝛿\deltaitalic_δでるた. However, it is possible to obtain alternative thickness estimates by relying on photon energy scans through the dopant resonance at low Q𝑄Qitalic_Q, provided that the two-dimensional dopant density N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT is known precisely. The arsenic L2,3subscript𝐿23L_{2,3}italic_L start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT-edge resonances are visible when measuring the reflected intensity as a function of X-ray energy at a fixed angle (see Fig. 3a), where the arsenic complex atomic scattering factor changes abruptly, while the silicon and oxygen atomic scattering factors are smooth functions of energy (see inset of Fig. 3a[38]. The scattering length density of the layers and the resonance intensity are determined by the effective three-dimensional dopant density N3D=N2D/δでるたsubscript𝑁3Dsubscript𝑁2D𝛿N_{\text{3D}}=N_{\text{2D}}/\deltaitalic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT / italic_δでるた and the Born equation Eq. (1), such that the thinner the dopant layer (δでるた𝛿\deltaitalic_δでるた), the stronger the resonance (because N3Dsubscript𝑁3DN_{\text{3D}}italic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT is larger) at any non-zero Q𝑄Qitalic_Q (described by Eq. (8)). The relative change in reflection, i.e., ΔでるたR/R=[R(1330eV)R(1320eV)]/R(1320eV),Δでるた𝑅𝑅delimited-[]𝑅1330eV𝑅1320eV𝑅1320eV\Delta R/R=[R(1330~{}\text{eV})-R(1320~{}\text{eV})]/R(1320~{}\text{eV}),roman_Δでるた italic_R / italic_R = [ italic_R ( 1330 eV ) - italic_R ( 1320 eV ) ] / italic_R ( 1320 eV ) , depends mostly on N3Dsubscript𝑁3DN_{\text{3D}}italic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT of the layer that undergoes the resonance, here the arsenic δでるた𝛿\deltaitalic_δでるた-layer, but its absolute value also hinges on the characteristics of all other layers [apparent from Eq. (6), see Methods]. As a consequence, if N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT and the layer depth are known, it is possible to obtain its thickness δでるた𝛿\deltaitalic_δでるた directly from the resonance spectrum.

The two-dimensional dopant density N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT is readily obtained by X-ray fluorescence [11]. In principle, it is straightforward to record the X-ray fluorescence and reflectivity simultaneously, as the fluorescence photons have an isotropic distribution. N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT was measured for one sample in this work (sample #1 shown in Fig. 2 & 3), and was found to be N2D=(2.77±0.14)×1014subscript𝑁2Dplus-or-minus2.770.14superscript1014N_{\text{2D}}=(2.77\pm 0.14)\times 10^{14}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = ( 2.77 ± 0.14 ) × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT cm-2 by the same method as in [11]. The relative change in reflection ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R at the arsenic L2,3subscript𝐿23L_{2,3}italic_L start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT-edges as a function of the dopant layer thickness was also calculated with the DYNA program [39] (see Fig. 3b, black). The layer depth used for this calculation was obtained in Fig. 2a. As expected, decreasing the dopant layer thickness and/or increasing N3Dsubscript𝑁3DN_{\text{3D}}italic_N start_POSTSUBSCRIPT 3D end_POSTSUBSCRIPT at fixed N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT, increases the relative change in resonance intensity ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R. In Fig. 3b, the experimental value of 0.09±0.01plus-or-minus0.090.010.09\pm 0.010.09 ± 0.01 for ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R extracted from Fig. 3a is shown in red. An arsenic δでるた𝛿\deltaitalic_δでるた-layer thickness of 1.6±0.5plus-or-minus1.60.51.6\pm 0.51.6 ± 0.5 nm is deduced. Within errors this agrees with the value from the previous section (δでるた=0.60.2+0.7𝛿superscriptsubscript0.60.20.7\delta=0.6_{-0.2}^{+0.7}italic_δでるた = 0.6 start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.7 end_POSTSUPERSCRIPT nm), but relies on fitting the resonance with each layer’s thickness, density, and roughness, resulting in a large number of fitting parameters, and prior knowledge about all layers in the sample.

Discussion

RCXR and SIMS experiments were performed on four samples containing an arsenic δでるた𝛿\deltaitalic_δでるた-layer. Figure 4 shows that the depths d𝑑ditalic_d of the shallower samples agree within error bars, whereas SIMS underestimates the depth of deeper Si:As δでるた𝛿\deltaitalic_δでるた-layers. This discrepancy might originate from the variability of the SIMS sputter rate during a measurement [44]. The δでるた𝛿\deltaitalic_δでるた-layer thickness (see Fig. 4a) is lower when measured with X-ray reflectometry, as is expected since the SIMS resolution is 2absent2\approx 2≈ 2 nm [45]. The values for the layer thickness measured with RCXR denote an upper bound, in our case determined by the maximum momentum transfer Qmaxsubscript𝑄maxQ_{\rm max}italic_Q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT=5 nm-1. In the present experiment where λらむだ𝜆\lambdaitalic_λらむだ=0.94 nm, Qmaxsubscript𝑄maxQ_{\rm max}italic_Q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT is fixed by θしーたmax=22subscript𝜃maxsuperscript22\theta_{\rm max}=22^{\circ}italic_θしーた start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 22 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, the maximum angle of incidence, which can easily be raised in experiments with experiments with next-generation synchrotron sources and instruments. Nevertheless, the RCXR results show that our arsenic δでるた𝛿\deltaitalic_δでるた-layer samples are as thin as 0.6 nm. Also, the upper-bound thicknesses measured here are in good agreement with results from ARPES [14], where Si:As samples fabricated in the same way were measured to be 0.4 to 0.7 nm thick (see shaded area in Fig. 4a) from the point of view of the 2D electron liquids hosted by the δでるた𝛿\deltaitalic_δでるた-layers. Additionally, one of the samples in this study (δでるたRCXR=1.4±0.5subscript𝛿𝑅𝐶𝑋𝑅plus-or-minus1.40.5\delta_{RCXR}=1.4\pm 0.5italic_δでるた start_POSTSUBSCRIPT italic_R italic_C italic_X italic_R end_POSTSUBSCRIPT = 1.4 ± 0.5 nm) was etched into a Hall bar geometry and contacted for low-temperature magneto-resistance [46], from which a conductive layer thickness of 0.97±0.02plus-or-minus0.970.020.97\pm 0.020.97 ± 0.02 nm was extracted [11]. Magneto-resistance of ten equivalent arsenic δでるた𝛿\deltaitalic_δでるた-layers yielded a thickness range from 0.4 nm to 1.8 nm (see shaded area in Fig. 4a), further corroborating our method and results.

Dopant δでるた𝛿\deltaitalic_δでるた-layers in silicon have been extensively studied with magneto-resistance [46], SIMS [8], and ARPES [14]. The agreement between data in Fig. 4 measured at a single energy above the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge (red) and by RCXR (black) shows that RCXR, treated within the first-order perturbation theory described here, is able to extract the arsenic dopant layer depth and thickness reliably without extensive structural modelling. The fact that the difference in phase shift of the oscillations on and off resonance matches the tabulated atomic phase shifts certifies the applicability of perturbation theory.

While RCXR is used here to enhance the sensitivity to a single two-dimensional dopant layer in silicon, naturally it also increases the sensitivity to specific layers in any multilayer material, including group IV and III-V semiconductors, as well as quantum materials where electron correlations are strong. For example, owing to the strong soft X-ray resonances of the transition metal L𝐿Litalic_L-edges and the oxygen K-edge, RCXR is ideally suited to non-destructively characterize oxide heterostructures [47, 48]. An ideal venue for our technique are the recently discovered superconducting infinite-layer nickelates [49, 50], for which the exact structure of the infinite-layer phase and other topotactically related phases are currently a matter of discussion [51, 52]. In particular, there is the possibility of an oxygen-ordered impurity superlattice [53], for which RCXR can be used to isolate and amplify the oxygen signal.

In conclusion, X-ray reflectometry can be made sensitive to specific elements in layered samples by performing a differential measurement above and below a resonance absorption edge of the respective element. We show that with this technique and a simple perturbation expansion of the differential reflectivity, it is possible to isolate the signal from one specific element without needing to model the material, yielding an upper bound as low as 0.9 nm for the thickness of arsenic-doped δでるた𝛿\deltaitalic_δでるた-layer buried in silicon with a 3D arsenic density <5%absentpercent5<5\%< 5 % of the total silicon density. In principle, the upper bound can be extended by increasing the X-ray fluence to scan to higher reflection angles, as well as by reducing the sample surface roughness. Combination of this approach with nano X-ray fluorescence detection [11] will make it possible to non-destructively characterize samples in three-dimensions in a single experiment. Finally, specular and off-specular RCXR with nano-X-ray beams featuring spot sizes of less than 10 nm [54] will enable imaging of patterned dopant structures such as the gates, sources and drains needed for future classical and quantum electronics.

Methods

First-order resonant-reflectometry theory. For a single interface between two materials, the reflectance R𝑅Ritalic_R for σしぐま𝜎\sigmaitalic_σしぐま polarised light, is given by the Fresnel equation

R=|n1sinθしーたin2sinθしーたtn1sinθしーたi+n2sinθしーたt|2,𝑅superscriptsubscript𝑛1subscript𝜃𝑖subscript𝑛2subscript𝜃𝑡subscript𝑛1subscript𝜃𝑖subscript𝑛2subscript𝜃𝑡2R=\left|\frac{n_{1}\sin\theta_{i}-n_{2}\sin\theta_{t}}{n_{1}\sin\theta_{i}+n_{% 2}\sin\theta_{t}}\right|^{2},italic_R = | divide start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_sin italic_θしーた start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_sin italic_θしーた start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_sin italic_θしーた start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_sin italic_θしーた start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (3)

where θしーたisubscript𝜃𝑖\theta_{i}italic_θしーた start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and θしーたtsubscript𝜃𝑡\theta_{t}italic_θしーた start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT are the beam incidence and transmission angles and are equal (θしーたi=θしーたt=θしーたsubscript𝜃𝑖subscript𝜃𝑡𝜃\theta_{i}=\theta_{t}=\thetaitalic_θしーた start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_θしーた start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_θしーた) for specular reflection as defined in Fig. 1. n1subscript𝑛1n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and n2subscript𝑛2n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the refractive indices of the top and bottom layer, respectively. The reflectance R𝑅Ritalic_R is related to the material’s atomic scattering factors f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT through the refractive index [35, 38]

n=1r02πぱいλらむだ2qNqfq=1λらむだ22πぱいρろー,𝑛1subscript𝑟02𝜋superscript𝜆2subscript𝑞subscript𝑁𝑞subscript𝑓𝑞1superscript𝜆22𝜋𝜌n=1-\frac{r_{0}}{2\pi}\lambda^{2}\sum_{q}N_{q}f_{q}=1-\frac{\lambda^{2}}{2\pi}\rho,italic_n = 1 - divide start_ARG italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_πぱい end_ARG italic_λらむだ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 1 - divide start_ARG italic_λらむだ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_πぱい end_ARG italic_ρろー , (4)

where r0subscript𝑟0r_{0}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the classical electron radius, λらむだ𝜆\lambdaitalic_λらむだ the photon wavelength, Nqsubscript𝑁𝑞N_{q}italic_N start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT and fq=f1,q+if2,qsubscript𝑓𝑞subscript𝑓1𝑞𝑖subscript𝑓2𝑞f_{q}=f_{1,q}+if_{2,q}italic_f start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 1 , italic_q end_POSTSUBSCRIPT + italic_i italic_f start_POSTSUBSCRIPT 2 , italic_q end_POSTSUBSCRIPT are the number of atoms per unit volume and the complex atomic scattering factor for an atom of element q𝑞qitalic_q, respectively. ρろー𝜌\rhoitalic_ρろー is the scattering length density. f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as well as f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and hence also ρろー𝜌\rhoitalic_ρろー, depend on the incident beam energy.

For multiple layers, the reflections at boundaries interfere. Depending on the incident beam angle θしーた𝜃\thetaitalic_θしーた and the layer thicknesses, this interference can be destructive or constructive. For two interfaces, as shown in Fig. 1, the condition for constructive interference is given by Bragg’s law 2dsinθしーた=mλらむだ2𝑑𝜃𝑚𝜆2d\sin\theta=m\lambda2 italic_d roman_sin italic_θしーた = italic_m italic_λらむだ, and the reflectance is periodic in the wave-vector change on reflection Q=4πぱいsinθしーた/λらむだ𝑄4𝜋𝜃𝜆Q=4\pi\sin\theta/\lambdaitalic_Q = 4 italic_πぱい roman_sin italic_θしーた / italic_λらむだ.

In most cases, rather than consisting of perfectly homogeneous and sharp layers, samples have a continuously varying scattering length density profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ). If scattering is weak, multiple scattering events can be neglected, and the reflected signal is simply the sum of the partial waves emanating from the different scatterers driven by the unperturbed incident field. The resulting far-field amplitude is given by the kinematic Born approximation [32, 33, 27]111The sign in the exponential in Eq. (5) depends on the convention for the imaginary part of ρろー𝜌\rhoitalic_ρろー. We follow Chantler [55], corresponding to the form of the Born approximation in Caticha [32].,

R(Q)=(4πぱい)2Q2|ρろー(z)eiQzdz|2=rF(Q)2|FQ(ρろー)|2,𝑅𝑄superscript4𝜋2superscript𝑄2superscript𝜌𝑧superscript𝑒𝑖𝑄𝑧differential-d𝑧2subscript𝑟𝐹superscript𝑄2superscriptsubscript𝐹𝑄𝜌2R(Q)=\frac{(4\pi)^{2}}{Q^{2}}\left|\int\rho(z)e^{-iQz}\ \mathrm{d}z\right|^{2}% =r_{F}(Q)^{2}|F_{Q}(\rho)|^{2},italic_R ( italic_Q ) = divide start_ARG ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG | ∫ italic_ρろー ( italic_z ) italic_e start_POSTSUPERSCRIPT - italic_i italic_Q italic_z end_POSTSUPERSCRIPT roman_d italic_z | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (5)

which is essentially the squared amplitude of the Fourier transform FQ(x)subscript𝐹𝑄𝑥F_{Q}(x)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_x ) of ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ) for a given Q𝑄Qitalic_Q-vector times the squared Fresnel reflectivity rF(Q)2=(4πぱい)2Q2subscript𝑟𝐹superscript𝑄2superscript4𝜋2superscript𝑄2r_{F}(Q)^{2}=\frac{(4\pi)^{2}}{Q^{2}}italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG ( 4 italic_πぱい ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_Q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. For a continuously varying density profile, the spatial resolution of reflectometry is of the order of the inverse of the greatest Q𝑄Qitalic_Q-vector, which is 0.2absent0.2\approx 0.2≈ 0.2 nm for the 1300 eV X-rays used in this work.

Data collected with X-ray reflectometry are commonly analysed using software that solves Maxwell’s equations throughout the material [56, 39]. To obtain good fits it is necessary to consider each layer’s atomic-species, density, thickness and roughness, resulting in a high number of fitting parameters. However, often one is faced with the problem of characterizing a layer with low contrast hosted by an otherwise known heterostructure, such as nano-electronic stacks involving III-V or group IV semiconductors, where dopant structures are typically buried, such as the silicon and oxide surface layers above our dopant δでるた𝛿\deltaitalic_δでるた-layers.

We introduce a new method to find and isolate the contribution to X-ray reflection from such a thin layer. First, we consider the general problem of a heterostructure with a scattering length density profile ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ), to which we add a second scattering length density profile δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ), accounting for replacement of atoms (e.g. As for silicon or aluminum for arsenic in III-V devices) from the original heterostructure. The Born approximation, expanding the square at the right of Eq. (5), yields three terms in the reflectivity of the perturbed heterostructure:

R(Q)rF(Q)2=|FQ(ρろー)|2+2[FQ(ρろー)FQ¯(δでるたρろー)]I(Q)+|FQ(δでるたρろー)|2.𝑅𝑄subscript𝑟𝐹superscript𝑄2superscriptsubscript𝐹𝑄𝜌2subscript2subscript𝐹𝑄𝜌¯subscript𝐹𝑄𝛿𝜌𝐼𝑄superscriptsubscript𝐹𝑄𝛿𝜌2\frac{R(Q)}{r_{F}(Q)^{2}}=|F_{Q}(\rho)|^{2}+\underbrace{2\Re\left[F_{Q}(\rho)% \bar{F_{Q}}(\delta\rho)\right]}_{I(Q)}+|F_{Q}(\delta\rho)|^{2}.divide start_ARG italic_R ( italic_Q ) end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + under⏟ start_ARG 2 roman_ℜ [ italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) over¯ start_ARG italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT end_ARG ( italic_δでるた italic_ρろー ) ] end_ARG start_POSTSUBSCRIPT italic_I ( italic_Q ) end_POSTSUBSCRIPT + | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_δでるた italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (6)

Eq. (6) contains a background from the unperturbed heterostructure, an interference term I(Q)𝐼𝑄I(Q)italic_I ( italic_Q ) which is first order in δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ), and a signal, of second order in δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ), from the perturbation. For a weak perturbation from a small number of atoms, we can ignore the second-order term, but the zeroth-order background remains troublesome as the major contributor to the reflectivity. However, if δでるたρろー(z)𝛿𝜌𝑧\delta\rho(z)italic_δでるた italic_ρろー ( italic_z ) could be modulated by a known prefactor without affecting ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ), differences of reflectivities for different prefactors would remove the unperturbed signal, leaving simply the interference term to analyze. The tunable photon energy of a synchrotron source provides a natural modulator for scattering lengths, especially near resonances which are unique fingerprints for particular elements. If the perturbation entails simply replacing atoms of species A with species B, then

δでるたρろー(z)=r0ΔでるたfN2Dδでるたρろー~(z),𝛿𝜌𝑧subscript𝑟0Δでるた𝑓subscript𝑁2D𝛿~𝜌𝑧\delta\rho(z)=r_{0}\Delta fN_{\text{2D}}\delta\tilde{\rho}(z),italic_δでるた italic_ρろー ( italic_z ) = italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_Δでるた italic_f italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT italic_δでるた over~ start_ARG italic_ρろー end_ARG ( italic_z ) , (7)

where Δでるたf=fBfAΔでるた𝑓subscript𝑓𝐵subscript𝑓𝐴\Delta f=f_{B}-f_{A}roman_Δでるた italic_f = italic_f start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT, N2Dδでるたρろー~(z)subscript𝑁2D𝛿~𝜌𝑧N_{\text{2D}}\delta\tilde{\rho}(z)italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT italic_δでるた over~ start_ARG italic_ρろー end_ARG ( italic_z ) is the 3D density of A atoms at depth z𝑧zitalic_z replaced by B atoms, N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT is the 2D density of the layer, and the integral over z𝑧zitalic_z of δでるたρろー~(z)𝛿~𝜌𝑧\delta\tilde{\rho}(z)italic_δでるた over~ start_ARG italic_ρろー end_ARG ( italic_z ) is unity. This integral is also the long wavelength (Q0𝑄0Q\to 0italic_Q → 0) limit of the Fourier transform of δでるたρろー~(z)𝛿~𝜌𝑧\delta\tilde{\rho}(z)italic_δでるた over~ start_ARG italic_ρろー end_ARG ( italic_z ), so that N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT sets the scale of the interference term in Eq. (6), which we can rewrite as

I(Q)=2r0N2D[Δでるたf¯FQ(ρろー)FQ¯(δでるたρろー~)].𝐼𝑄2subscript𝑟0subscript𝑁2D¯Δでるた𝑓subscript𝐹𝑄𝜌¯subscript𝐹𝑄𝛿~𝜌I(Q)=2r_{0}N_{\text{2D}}\Re\left[\bar{\Delta f}F_{Q}(\rho)\bar{F_{Q}}(\delta% \tilde{\rho})\right].italic_I ( italic_Q ) = 2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT roman_ℜ [ over¯ start_ARG roman_Δでるた italic_f end_ARG italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) over¯ start_ARG italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT end_ARG ( italic_δでるた over~ start_ARG italic_ρろー end_ARG ) ] . (8)

To progress, we assume that for a relatively deep δでるた𝛿\deltaitalic_δでるた-layer at depth d𝑑ditalic_d, δでるたρろー~(z)=h(zd)𝛿~𝜌𝑧𝑧𝑑\delta\tilde{\rho}(z)=h(z-d)italic_δでるた over~ start_ARG italic_ρろー end_ARG ( italic_z ) = italic_h ( italic_z - italic_d ), where h(z)𝑧h(z)italic_h ( italic_z ) is an even function. Hence FQ(δでるたρろー~)=exp(iQd)H(Q)subscript𝐹𝑄𝛿~𝜌𝑖𝑄𝑑𝐻𝑄F_{Q}(\delta\tilde{\rho})=\exp(-iQd)H(Q)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_δでるた over~ start_ARG italic_ρろー end_ARG ) = roman_exp ( - italic_i italic_Q italic_d ) italic_H ( italic_Q ) where H(Q)𝐻𝑄H(Q)italic_H ( italic_Q ) is real, giving

I(Q)=2r0N2DH(Q)[Δでるたf¯exp(iQd)FQ(ρろー)].𝐼𝑄2subscript𝑟0subscript𝑁2D𝐻𝑄¯Δでるた𝑓𝑖𝑄𝑑subscript𝐹𝑄𝜌I(Q)=2r_{0}N_{\text{2D}}H(Q)\Re\left[\bar{\Delta f}\exp(iQd)F_{Q}(\rho)\right].italic_I ( italic_Q ) = 2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT italic_H ( italic_Q ) roman_ℜ [ over¯ start_ARG roman_Δでるた italic_f end_ARG roman_exp ( italic_i italic_Q italic_d ) italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) ] . (9)

If d𝑑ditalic_d is the largest characteristic length in the problem, i.e. d𝑑ditalic_d is much larger than the thickness of a probable surface oxide layer which contributes to FQ(ρろー)subscript𝐹𝑄𝜌F_{Q}(\rho)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ), Eq. (9) gives several results:

  1. 1.

    There are rapid oscillations with period 2πぱい/d2𝜋𝑑2\pi/d2 italic_πぱい / italic_d and a phase fixed largely by the complex scattering amplitude ΔでるたfΔでるた𝑓\Delta froman_Δでるた italic_f, implying that changes in photon energy will change the phase of the oscillations. On the other hand, changes in N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT or H(Q)𝐻𝑄H(Q)italic_H ( italic_Q ) (as long as h(z)𝑧h(z)italic_h ( italic_z ) remains even) will not change this phase. Furthermore, the third term in Eq. (6) scales as |Δでるたf|2superscriptΔでるた𝑓2|\Delta f|^{2}| roman_Δでるた italic_f | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and, therefore, does not undergo a phase shift on changing photon energy. This means that the importance of the interference term in Eq. (6) and the validity of first-order perturbation theory is readily checked by examining the phase shift.

  2. 2.

    Oscillation amplitudes scale by N2D|ΔでるたfFQ(ρろー)|H(Q)subscript𝑁2DΔでるた𝑓subscript𝐹𝑄𝜌𝐻𝑄N_{\text{2D}}\left|\Delta fF_{Q}(\rho)\right|H(Q)italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT | roman_Δでるた italic_f italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | italic_H ( italic_Q ), meaning that if we operate near a resonance of the substitute atom B but far from resonances of the unperturbed heterostructure, then there will be an anomaly at the resonance energy for B.

  3. 3.

    For an infinitesimally thin δでるた𝛿\deltaitalic_δでるた-layer, h(z)=δでるた(z)𝑧𝛿𝑧h(z)=\delta(z)italic_h ( italic_z ) = italic_δでるた ( italic_z ) and H(Q)=1𝐻𝑄1H(Q)=1italic_H ( italic_Q ) = 1, the maximum amplitude of the anomaly is 2r0N2D|ΔでるたfFQ(ρろー)|2subscript𝑟0subscript𝑁2DΔでるた𝑓subscript𝐹𝑄𝜌2r_{0}N_{\text{2D}}\left|\Delta fF_{Q}(\rho)\right|2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT | roman_Δでるた italic_f italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | independent of the Q𝑄Qitalic_Q sampled. For a Gaussian, h(z)=1/(2πぱいσしぐま)exp([z/σしぐま]2/2)𝑧12𝜋𝜎superscriptdelimited-[]𝑧𝜎22h(z)=1/\left(\sqrt{2\pi}\sigma\right)\exp(-[z/\sigma]^{2}/2)italic_h ( italic_z ) = 1 / ( square-root start_ARG 2 italic_πぱい end_ARG italic_σしぐま ) roman_exp ( - [ italic_z / italic_σしぐま ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ), so that H(Q)=exp([Qσしぐま]2/2)𝐻𝑄superscriptdelimited-[]𝑄𝜎22H(Q)=\exp(-[Q\sigma]^{2}/2)italic_H ( italic_Q ) = roman_exp ( - [ italic_Q italic_σしぐま ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 ), implying that if N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT and |FQ(ρろー)|subscript𝐹𝑄𝜌|F_{Q}(\rho)|| italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | are known by other means, the amplitude of the anomaly at any non-zero Q𝑄Qitalic_Q will be reduced by the factor H(Q)𝐻𝑄H(Q)italic_H ( italic_Q ), from which the δでるた𝛿\deltaitalic_δでるた-layer thickness parameter δでるた=22πぱいσしぐま𝛿22𝜋𝜎\delta=2\sqrt{2}\pi\sigmaitalic_δでるた = 2 square-root start_ARG 2 end_ARG italic_πぱい italic_σしぐま can be extracted.

  4. 4.

    Assuming that FQ(ρろー)subscript𝐹𝑄𝜌F_{Q}(\rho)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) and H(Q)𝐻𝑄H(Q)italic_H ( italic_Q ) vary slowly with Q𝑄Qitalic_Q on the scale of the oscillations from the δでるた𝛿\deltaitalic_δでるた-layer, a moving average or low-pass filter with window of width 2πぱい/d2𝜋𝑑2\pi/d2 italic_πぱい / italic_d will remove the interference term in Eq. (6), leaving an estimate |F|LF2subscriptsuperscript𝐹2LF|F|^{2}_{\mathrm{LF}}| italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT of the strong zeroth-order term |FQ(ρろー)|2superscriptsubscript𝐹𝑄𝜌2|F_{Q}(\rho)|^{2}| italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The interference term is then estimated as

    I(Q)R(Q)rF(Q)2|F|LF2.𝐼𝑄𝑅𝑄subscript𝑟𝐹superscript𝑄2subscriptsuperscript𝐹2LFI(Q)\approx\frac{R(Q)}{r_{F}(Q)^{2}}-|F|^{2}_{\mathrm{LF}}.italic_I ( italic_Q ) ≈ divide start_ARG italic_R ( italic_Q ) end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT . (10)

    Thus we can approximate δでるたρろー𝛿𝜌\delta{\rho}italic_δでるた italic_ρろー as the inverse Fourier transform of the measured oscillations divided by |F|LF2|FQ(ρろー)|2subscriptsuperscript𝐹2LFsuperscriptsubscript𝐹𝑄𝜌2\sqrt{|F|^{2}_{\mathrm{LF}}}\approx|F_{Q}(\rho)|^{2}square-root start_ARG | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT end_ARG ≈ | italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, assuming that FQ(ρろー)subscript𝐹𝑄𝜌F_{Q}(\rho)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) is imaginary for non-zero Q𝑄Qitalic_Q, which will be the case if ρろー(z)𝜌𝑧\rho(z)italic_ρろー ( italic_z ) is a step function at z=0𝑧0z=0italic_z = 0, a reasonable approximation for samples of the type considered here:

    δでるたρろー¯=FT1(I(Q)2|F|LF2).¯𝛿𝜌superscriptFT1𝐼𝑄2subscriptsuperscript𝐹2LF\bar{\delta{\rho}}=\mathrm{FT}^{-1}\left(\frac{I(Q)}{2\sqrt{|F|^{2}_{\mathrm{% LF}}}}\right).over¯ start_ARG italic_δでるた italic_ρろー end_ARG = roman_FT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG italic_I ( italic_Q ) end_ARG start_ARG 2 square-root start_ARG | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT end_ARG end_ARG ) . (11)

    Finally, for the Gaussian δでるた𝛿\deltaitalic_δでるた-layer profile discussed under point 3., dividing Eq. (9) by |F|LF2subscriptsuperscript𝐹2LF\sqrt{|F|^{2}_{\mathrm{LF}}}square-root start_ARG | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT end_ARG gives:

    I(Q)|F|LF2=2r0N2D|Δでるたf|exp((Qσしぐま)22)cos(Qdϕ),𝐼𝑄subscriptsuperscript𝐹2LF2subscript𝑟0subscript𝑁2DΔでるた𝑓superscript𝑄𝜎22𝑄𝑑italic-ϕ\frac{I(Q)}{\sqrt{|F|^{2}_{\mathrm{LF}}}}=2r_{0}N_{\text{2D}}|\Delta f|\exp% \left(-\frac{(Q\sigma)^{2}}{2}\right)\cos(Qd-\phi),divide start_ARG italic_I ( italic_Q ) end_ARG start_ARG square-root start_ARG | italic_F | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_LF end_POSTSUBSCRIPT end_ARG end_ARG = 2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT | roman_Δでるた italic_f | roman_exp ( - divide start_ARG ( italic_Q italic_σしぐま ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ) roman_cos ( italic_Q italic_d - italic_ϕ ) , (12)

    where ϕ=arg(Δでるたf)italic-ϕargΔでるた𝑓\phi=\text{arg}(\Delta f)italic_ϕ = arg ( roman_Δでるた italic_f ).

Instead of directly analyzing data as described above or via modelling software, it is also useful to consider differences ΔでるたR(Q)=R(E,Q)R(E,Q)Δでるた𝑅𝑄𝑅superscript𝐸𝑄𝑅𝐸𝑄\Delta R(Q)=R(E^{\prime},Q)-R(E,Q)roman_Δでるた italic_R ( italic_Q ) = italic_R ( italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_Q ) - italic_R ( italic_E , italic_Q ) for two different X-ray energies E𝐸Eitalic_E and Esuperscript𝐸E^{\prime}italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT close to a type B atom resonance, such that changes in terms not containing B atom contributions can be omitted. In terms of the Born approximation this leads to:

ΔでるたR(Q)rF(Q)2=2[FQ(ρろー)(FQE¯(δでるたρろー~)FQE¯(δでるたρろー~))],Δでるた𝑅𝑄subscript𝑟𝐹superscript𝑄22subscript𝐹𝑄𝜌¯superscriptsubscript𝐹𝑄superscript𝐸𝛿~𝜌¯superscriptsubscript𝐹𝑄𝐸𝛿~𝜌\displaystyle\frac{\Delta R(Q)}{r_{F}(Q)^{2}}=2\Re\left[F_{Q}(\rho)\left(\bar{% F_{Q}^{E^{\prime}}}(\delta\tilde{\rho})-\bar{F_{Q}^{E}}(\delta\tilde{\rho})% \right)\right],divide start_ARG roman_Δでるた italic_R ( italic_Q ) end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 2 roman_ℜ [ italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) ( over¯ start_ARG italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG ( italic_δでるた over~ start_ARG italic_ρろー end_ARG ) - over¯ start_ARG italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_E end_POSTSUPERSCRIPT end_ARG ( italic_δでるた over~ start_ARG italic_ρろー end_ARG ) ) ] , (13)

where FQE(ρろー)=FQE(ρろー)=FQ(ρろー)superscriptsubscript𝐹𝑄superscript𝐸𝜌superscriptsubscript𝐹𝑄𝐸𝜌subscript𝐹𝑄𝜌F_{Q}^{E^{\prime}}(\rho)=F_{Q}^{E}(\rho)=F_{Q}(\rho)italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_ρろー ) = italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_E end_POSTSUPERSCRIPT ( italic_ρろー ) = italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) as it does not contain type B atoms. This shows that measuring the difference in reflectometry at two energies near a δでるた𝛿\deltaitalic_δでるた-layer’s atomic absorption edge isolates the dopant’s signal, dependent to first order only on the change in the dopant’s atomic scattering factors at the absorption edge.

Applying the assumptions that led to Eq. (9), we can recast Eq. (13) as

ΔでるたR(Q)rF(Q)2=2r0N2DH(Q)[(ΔでるたfE¯ΔでるたfE¯)exp(iQd)FQ(ρろー)].Δでるた𝑅𝑄subscript𝑟𝐹superscript𝑄22subscript𝑟0subscript𝑁2D𝐻𝑄¯Δでるたsubscript𝑓superscript𝐸¯Δでるたsubscript𝑓𝐸𝑖𝑄𝑑subscript𝐹𝑄𝜌\frac{\Delta R(Q)}{r_{F}(Q)^{2}}=2r_{0}N_{\text{2D}}H(Q)\Re\left[\left(\bar{% \Delta f_{E^{\prime}}}-\bar{\Delta f_{E}}\right)\exp(iQd)F_{Q}(\rho)\right].divide start_ARG roman_Δでるた italic_R ( italic_Q ) end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 2 italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT italic_H ( italic_Q ) roman_ℜ [ ( over¯ start_ARG roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT end_ARG - over¯ start_ARG roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_ARG ) roman_exp ( italic_i italic_Q italic_d ) italic_F start_POSTSUBSCRIPT italic_Q end_POSTSUBSCRIPT ( italic_ρろー ) ] . (14)

This matches Eq. (9), with I(Q)ΔでるたR(Q)/rF(Q)2𝐼𝑄Δでるた𝑅𝑄subscript𝑟𝐹superscript𝑄2I(Q)\to\Delta R(Q)/r_{F}(Q)^{2}italic_I ( italic_Q ) → roman_Δでるた italic_R ( italic_Q ) / italic_r start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_Q ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and ΔでるたfΔでるたfEΔでるたfEΔでるた𝑓Δでるたsubscript𝑓superscript𝐸Δでるたsubscript𝑓𝐸\Delta f\to\Delta f_{E^{\prime}}-\Delta f_{E}roman_Δでるた italic_f → roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT - roman_Δでるた italic_f start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT, so Eq. (11) & (12) apply with these substitutions. This resonant contrast analysis has the advantage of isolating the interference term [Eq. (13)] without needing to separate signals at different frequencies, as required when analysing reflectometry at a single energy [Eq. (10)].

Sample preparation. Four 2×9292\times 92 × 9 mm2 Si(001) samples were diced from Czochralski-grown wafers (bulk doping <5×1014absent5superscript1014<5\times 10^{14}< 5 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT cm-3), and cleaned ultrasonically in acetone, followed by isopropyl alcohol. Each sample was thermally outgassed in vacuum (base pressure <5×1010absent5superscript1010<5\times 10^{-10}< 5 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT mbar) for > 8 h at 600 C, then flash annealed multiple times at 1200 C using direct current resistive sample heating. This temperature was monitored using an infrared pyrometer (IMPAC IGA50-LO plus) with an uncertainty of ±30\pm 30~{}^{\circ}± 30 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPTC. Each sample was dosed with AsH3, to the saturation As density of (1.6±0.3)×1014plus-or-minus1.60.3superscript1014(1.6\pm 0.3)\times 10^{14}( 1.6 ± 0.3 ) × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT cm-2 [5], and heated to 350 C for 12121-21 - 2 min to incorporate the arsenic into the silicon lattice [57]. Afterwards, a 14141-41 - 4 nm silicon “locking layer” was deposited, without resistive sample heating, to confine the arsenic [58, 5]. Three of the samples were then heated to 500 C for 15 s, to improve electrical activation, whereas sample #1 was not heated. More silicon (1471147114-7114 - 71 nm depending on sample) was deposited, with the sample at 250 C. Silicon was deposited at a rate of 0.10.40.10.40.1-0.40.1 - 0.4 nm/min using a silicon solid sublimation source (SUSI-40, MBE Komponenten GmbH). During deposition, the sample temperature was indirectly monitored by the sample resistance, while heating using a direct current resistive sample heater. A control sample with a buried oxide layer was also grown. The process was as above, except without the flash anneal or AsH3 dose, and the silicon deposition (15 nm) was done without sample heating. All samples and parameters are shown in Tab.  1. The sample in Fig. 2 and 3 was etched with HF prior to the RCXR measurement to make the surface oxide layer thinner, which suppresses an additional slow oscillation to the reflectometry.

# Dopant Locking layer 500 C anneal dSIMSsubscript𝑑SIMSd_{\text{SIMS}}italic_d start_POSTSUBSCRIPT SIMS end_POSTSUBSCRIPT (nm) dXRRsubscript𝑑XRRd_{\text{XRR}}italic_d start_POSTSUBSCRIPT XRR end_POSTSUBSCRIPT (nm) δでるたXRRsubscript𝛿XRR\delta_{\text{XRR}}italic_δでるた start_POSTSUBSCRIPT XRR end_POSTSUBSCRIPT (nm)
1 As Yes No 16.0±plus-or-minus\pm±1.7 16.8±plus-or-minus\pm±1.7 0.8 (0.61.2)
2 As Yes Yes 15.3±plus-or-minus\pm±1.7 17.9±plus-or-minus\pm±1.7 0.6 (0.41.3)
3 As Yes Yes 28.7±plus-or-minus\pm±2.3 32±plus-or-minus\pm±2 1.4 (1.02.0)
4 As Yes Yes 70.5±plus-or-minus\pm±3.0 77±plus-or-minus\pm±3 1.2 (1.01.4)
5 SiO2 No No 13.4±plus-or-minus\pm±1.8 15.3±plus-or-minus\pm±0.8
Table 1: Samples details, depths d𝑑ditalic_d and thicknesses δでるた𝛿\deltaitalic_δでるた measured by SIMS and XRR. SIMS errors are from the uncertainty in the sputter rate. XRR depth errors are from the FWHM of the main peak in the signal’s Fourier transform. The uncertainty on the thicknesses is the 95% confidence interval of fits to Eq. (2).
Refer to caption
Figure 5: Arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge fits. Black dots correspond to the reflectometry data shown in Fig. 3a. The lines are fits to the data with colours corresponding to a Si:As layer thickness δでるた𝛿\deltaitalic_δでるた as indicated in the legend.

X-ray scattering. RCXR was performed at the RESOXS endstation of the SIM beamline at the Swiss Light Source synchrotron of the Paul Scherrer Institute [59, 60]. Samples were kept in high vacuum at 10-8 mbar, and at room temperature. They were mounted on a rotatable holder such that the incidence angle θしーた𝜃\thetaitalic_θしーた could be swept from 0superscript00^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT to 90superscript9090^{\circ}90 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. The beam energy was set between 1200 and 1400 eV with linear polarization and a spot size set between 500 and 120 µm in the horizontal and 25 to 50 µm in the vertical direction.

Arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge reflectometry simulations. To assess the expected change in reflection ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R at the arsenic L3subscript𝐿3L_{3}italic_L start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-edge, we use the DYNA program [39] to obtain the best fit to the resonant spectrum for different Si:As layer thicknesses. The resulting fits are shown in Fig. 5, each taking into account the oxide, silicon, and Si:As layer thicknesses, densities, and roughness. The silicon and oxide density is known, and the density of the Si:As layer is calculated from the 2D density N2Dsubscript𝑁2DN_{\text{2D}}italic_N start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT measured by X-ray fluorescence and the layer’s thickness δでるた𝛿\deltaitalic_δでるた (N3D=N2D/δでるたsubscript𝑁3Dsubscript𝑁2D𝛿N_{\mathrm{3D}}=N_{\mathrm{2D}}/\deltaitalic_N start_POSTSUBSCRIPT 3 roman_D end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT 2 roman_D end_POSTSUBSCRIPT / italic_δでるた), such that the simulated ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R captures the larger arsenic density in thinner Si:As δでるた𝛿\deltaitalic_δでるた-layers. The surface roughness obtained from the fits varies in the range 0.12Å0.12Å0.1-2~{}\text{\AA}0.1 - 2 Å, in agreement with STM measurements where roughness was found to be in the range of 1 Å. The experimental values for R(1320eV)𝑅1320eVR(1320~{}\text{eV})italic_R ( 1320 eV ) and R(1330eV)𝑅1330eVR(1330~{}\text{eV})italic_R ( 1330 eV ) are taken as the mean of the data from 1312 to 1322 eV and 1330 to 1340 eV, respectively. The standard deviation is determined in the same intervals, yielding ΔでるたR/R=0.09±0.01Δでるた𝑅𝑅plus-or-minus0.090.01\Delta R/R=0.09\pm 0.01roman_Δでるた italic_R / italic_R = 0.09 ± 0.01. ΔでるたRΔでるた𝑅\Delta Rroman_Δでるた italic_R of the simulation is taken to be [max(R)min(R)]delimited-[]max𝑅min𝑅[\text{max}(R)-\text{min}(R)][ max ( italic_R ) - min ( italic_R ) ] in the interval 1320 to 1340 eV. The values for ΔでるたR/RΔでるた𝑅𝑅\Delta R/Rroman_Δでるた italic_R / italic_R from the fits are shown as black dots on Fig. 3b with a “shape preserving interpolant” line as a guide to the eye.

Secondary ion mass spectrometry. Time-of-flight SIMS (IONTOF ToF-SIMS5) was conducted on all samples with a 25 keV, 1 pA Bi+ primary ion beam in high current bunch mode, and a 500 eV, 35–50 nA Cs+ sputter beam. Depth profiles were made with a 300 ×\times× 300 Î¼m2 sputter crater, within which the analytical region was the central 50 ×\times× 50 Î¼m2 or 100 ×\times× 100 Î¼m2. The sputter rate was determined by measuring the crater depth with an interference microscope (Zygo NewView NX2). The crater depth was measured from line profiles of the topography in different directions, where the uncertainty was estimated from the standard deviation of these measurements. The results are shown in Fig. 6.

Refer to caption
Figure 6: SIMS depth profiles. Arsenic depth profiles of the four Si:As δでるた𝛿\deltaitalic_δでるた-layers of varying depth, along with the oxide profile of the buried oxide layer (see Tab. 1).

Scanning probe microscopy. STM was conducted using an Omicron variable temperature system. AFM was preformed with a Bruker Dimension Icon with a ScanAsyst-Air cantilever, using peak force tapping mode.

Acknowledgements.
We acknowledge the Paul Scherrer Institute, Villigen, Switzerland for provision of synchrotron radiation at the RESOXS endstation at the SIM beamline and the microXAS beamline of the Swiss Light Source. This project received funding from the European Research Council under the European Union’s Horizon 2020 research and innovation program, within the Hidden, Entangled and Resonating Order (HERO) project with Grant Agreement 810451. The project was financially supported by the Engineering and Physical Sciences Research Council (EPSRC) Grant Numbers EP/M009564/1, EP/R034540/1, EP/V027700/1, and EP/W000520/1, as well as, Innovate UK Grant Number UKRI/75574. N.D. was partially supported by Swiss National Science Foundation Contract 175867. E.S. received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie Grant Agreement 884104 (PSI-FELLOW-III-3i). H.U. was supported by the National Centers of Competence in Research in Molecular Ultrafast Science and Technology (Grant Number 51NF40-183615) from the Swiss National Science Foundation, and the European Union’s Horizon 2020 research and innovation program Marie Sklodowska-Curie Grant Agreement 801459 (FP-RESOMUS). J.B, K.S., and P.C.C. were supported by the EPSRC Centre for Doctoral Training in Advanced Characterization of Materials (Grant Number EP/L015277/1), and by the Paul Scherrer Institute.

References

  • Stroscio and Eigler [1991] J. A. Stroscio and D. M. Eigler, Atomic and molecular manipulation with the scanning tunneling microscope, Science 254, 1319 (1991).
  • Crommie et al. [1993] M. F. Crommie, C. P. Lutz, and D. M. Eigler, Confinement of electrons to quantum corrals on a metal surface, Science 262, 218 (1993).
  • Ruess et al. [2004] F. J. Ruess et al., Toward atomic-scale device fabrication in silicon using scanning probe microscopy, Nano Lett. 4, 1969 (2004).
  • Fuechsle et al. [2012] M. Fuechsle et al., Single-atom transistor, Nat. Nanotechnol. 7, 242 (2012).
  • Stock et al. [2020] T. J. Z. Stock et al., Atomic-scale patterning of arsenic in silicon by scanning tunneling microscopy, ACS Nano 14, 3316 (2020).
  • Dwyer et al. [2021] K. J. Dwyer et al., B-doped δでるた𝛿\deltaitalic_δでるた-layers and nanowires from area-selective deposition of BCl3subscriptBCl3\rm BCl_{3}roman_BCl start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT on Si(100)Si100\rm Si(100)roman_Si ( 100 )ACS Appl. Mater. Interfaces 13, 41275 (2021).
  • Jeong et al. [2018] W. C. Jeong et al., True 7 nm Platform Technology featuring Smallest FinFET and Smallest SRAM cell by EUV, Special Constructs and 3rd Generation Single Diffusion Break, in 2018 IEEE Symposium on VLSI Technology (2018) pp. 59–60.
  • Wang et al. [2018] X. Wang et al., Quantifying atom-scale dopant movement and electrical activation in Si:P monolayers, Nanoscale 10, 4488 (2018).
  • Chang and Lauhon [2018] A. S. Chang and L. J. Lauhon, Atom probe tomography of nanoscale architectures in functional materials for electronic and photonic applications, Curr. Opin. Solid State Mater. Sci. 22, 171 (2018).
  • Werner and Boudewijn [1984] H. W. Werner and P. R. Boudewijn, A comparison of SIMS with other techniques based on ion-beam solid interactions, Vacuum 34, 83 (1984).
  • D’Anna et al. [2023] N. D’Anna et al., Non-destructive X-ray imaging of patterned delta-layer devices in silicon, Adv. Electron. Mater. 9, 2201212 (2023).
  • Masteghin et al. [2024] M. G. Masteghin et al., Benchmarking of X-Ray fluorescence microscopy with ion beam implanted samples showing detection sensitivity of hundreds of atoms, Small Methods 2024, 2301610 (2024).
  • Schülli and Leake [2018] T. U. Schülli and S. J. Leake, X-ray nanobeam diffraction imaging of materials, Curr. Opin. Solid State Mater. Sci. 22, 188 (2018).
  • Constantinou et al. [2023] P. Constantinou et al., Momentum-space imaging of ultra-thin electron liquids in δでるた𝛿\deltaitalic_δでるた-doped silicon, Adv. Sci. 10, 2302101 (2023).
  • Katzenmeyer et al. [2020] A. M. Katzenmeyer et al., Assessing atomically thin delta-doping of silicon using mid-infrared ellipsometry, J. Mater. Res. 35, 2098 (2020).
  • Young et al. [2023] S. M. Young et al., Suppression of midinfrared plasma resonance due to quantum confinement in δでるた𝛿\deltaitalic_δでるた-doped silicon, Phys. Rev. Appl. 20, 024043 (2023).
  • Barber et al. [2022] M. E. Barber, E. Y. Ma, and Z.-X. Shen, Microwave impedance microscopy and its application to quantum materials, Nat. Rev. Phys. 4, 61 (2022).
  • Gramse et al. [2017] G. Gramse et al., Nondestructive imaging of atomically thin nanostructures buried in silicon, Sci. Adv. 3, e1602586 (2017).
  • Gramse et al. [2020] G. Gramse et al., Nanoscale imaging of mobile carriers and trapped charges in delta doped silicon p–n junctions, Nat. Electron. 3, 531 (2020).
  • Ng et al. [2020] K. S. H. Ng et al., Scanned single-electron probe inside a silicon electronic device, ACS Nano 14, 9449 (2020).
  • Amsel and Lanford [1984] G. Amsel and W. A. Lanford, Nuclear reaction techniques in materials analysis, Annu. Rev. Nucl. Part. Sci. 34 (1984).
  • Goncharova [2020] L. V. Goncharova, Rutherford backscattering spectroscopy (RBS) and medium energy ion scattering (MEIS), in Characterization of Nanoparticles (Elsevier, 2020) pp. 441–455.
  • Trombini et al. [2019] H. Trombini et al., Profiling As plasma doped Si/SiO2 with molecular ions, Thin Solid Films 692, 137536 (2019).
  • Sanchez et al. [2011] D. Sanchez et al., Structural characterization of Pb nanoislands in SiO2/Si interface synthesized by ion implantation through MEIS analysis, Surf. Sci. 605, 654 (2011).
  • Holler et al. [2017] M. Holler et al., High-resolution non-destructive three-dimensional imaging of integrated circuits, Nature 543, 402 (2017).
  • de Jonge et al. [2010] M. D. de Jonge et al., Quantitative 3D elemental microtomography of Cyclotella meneghiniana at 400-nm resolution, Proc. Natl. Acad. Sci. U.S.A. 107, 15676 (2010).
  • Zhou and Chen [1995] X.-L. Zhou and S.-H. Chen, Theoretical foundation of X-ray and neutron reflectometry, Phys. Rep. 257, 223 (1995).
  • Slijkerman et al. [1990] W. F. J. Slijkerman et al., X‐ray reflectivity of an Sb delta‐doping layer in silicon, J. Appl. Phys. 68, 5105 (1990).
  • Macke et al. [2014] S. Macke et al., Element specific monolayer depth profiling, Adv. Mater. 26, 6554 (2014).
  • Parratt [1954] L. G. Parratt, Surface studies of solids by total reflection of X-rays, Phys. Rev. 95, 359 (1954).
  • Tolan and Press [1998] M. Tolan and W. Press, X-ray and neutron reflectivity, Z. Kristallogr. - Cryst. Mater. 213, 319 (1998).
  • Caticha [1995] A. Caticha, Reflection and transmission of X-rays by graded interfaces, Phys. Rev. B 52, 9214 (1995).
  • Als-Nielsen and McMorrow [2011] J. Als-Nielsen and D. McMorrow, Refraction and reflection from interfaces, in Elements of Modern X‐ray Physics (Wiley, 2011) pp. 147–205.
  • Al-Bayati et al. [1991] A. H. Al-Bayati, K. G. Orrman-Rossiter, J. A. van den Berg, and D. G. Armour, Composition and structure of the native Si oxide by high depth resolution medium energy ion scatering, Surf. Sci. 241, 91 (1991).
  • Henke et al. [1993] B. L. Henke, E. M. Gullikson, and J. C. Davis, X-ray interactions: Photoabsorption, scattering, transmission, and reflection at E=5030,000𝐸5030000{E}=50-30,000italic_E = 50 - 30 , 000 eV, Z=192𝑍192{Z}=1-92italic_Z = 1 - 92At. Data Nucl. Data Tables 54, 181 (1993).
  • Nevot and Croce [1980] L. Nevot and P. Croce, Characterization of surfaces by grazing X-ray reflection–application to the study of polishing of some silicate glasses, Rev. Phys. Appl. 15, 761 (1980).
  • Goh and Simmons [2009] K. E. J. Goh and M. Y. Simmons, Impact of Si growth rate on coherent electron transport in Si:P delta-doped devices, Appl. Phys. Lett. 95, 142104 (2009).
  • Chantler et al. [2005] C. T. Chantler et al., X-ray form factor, attenuation and scattering tables (version 2.1), 10.18434/T4HS32 (2005).
  • Elzo et al. [2012] M. Elzo et al., X-ray resonant magnetic reflectivity of stratified magnetic structures: Eigenwave formalism and application to a W/Fe/W trilayer, J. Magn. Magn. Mater. 324, 105 (2012).
  • Liu et al. [2020] Y. Liu et al., Coherent epitaxial semiconductor–ferromagnetic insulator InAs/EuS interfaces: Band alignment and magnetic structure, ACS Appl. Mater. Interfaces 12, 8780 (2020).
  • Ilse et al. [2023] S. E. Ilse, G. Schütz, and E. Goering, Voltage X-ray reflectometry: A method to study electric-field-induced changes in interfacial electronic structures, Phys. Rev. Lett. 131, 036201 (2023).
  • Su et al. [2012] H.-C. Su, C.-H. Lee, M.-Z. Lin, and T.-W. Huang, A comparison between X-ray reflectivity and atomic force microscopy on the characterization of a surface roughness, Chin. J. Phys. 50, 291 (2012).
  • Freitag and Clemens [2001] J. M. Freitag and B. M. Clemens, Nonspecular X-ray reflectivity study of roughness scaling in Si/Mo multilayers, J. Appl. Phys. 89, 1101 (2001).
  • Homma et al. [2003] Y. Homma et al., Evaluation of the sputtering rate variation in SIMS ultra-shallow depth profiling using multiple short-period delta layers, Surf. Interface Anal. 35, 544 (2003).
  • Hagmann et al. [2020] J. A. Hagmann et al., Electron-electron interactions in low-dimensional Si:P delta layers, Phys. Rev. B 101, 245419 (2020).
  • Hagmann et al. [2018] J. A. Hagmann et al., High resolution thickness measurements of ultrathin Si:P monolayers using weak localization, Appl. Phys. Lett. 112, 043102 (2018).
  • Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, From quantum matter to high-temperature superconductivity in copper oxides, Nature 518, 179 (2015).
  • Logvenov et al. [2009] G. Logvenov, A. Gozar, and I. Bozovic, High-temperature superconductivity in a single copper-oxygen plane, Science 326, 699 (2009).
  • Li et al. [2019] D. Li et al., Superconductivity in an infinite-layer nickelate, Nature 572, 624 (2019).
  • Zeng et al. [2022] S. Zeng et al., Superconductivity in infinite-layer nickelate La1-xCaxNiO2 thin films, Sci. Adv. 8, eabl9927 (2022).
  • Lee et al. [2020] K. Lee et al., Aspects of the synthesis of thin film superconducting infinite-layer nickelates, APL Mater. 8, 041107 (2020).
  • Nomura and Arita [2022] Y. Nomura and R. Arita, Superconductivity in infinite-layer nickelates, Rep. Prog. Phys. 85, 052501 (2022).
  • Parzyck et al. [2024] C. T. Parzyck et al., Absence of 3a03subscript𝑎03a_{0}3 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT charge density wave order in the infinite-layer nickelate NdNiO2Nat. Mater. 23, 486 (2024).
  • Villar et al. [2018] F. Villar et al., Nanopositioning for the ESRF ID16A nano-imaging beamline, Synchrotron Radiat. News 31, 9 (2018).
  • Chantler [2000] C. T. Chantler, Detailed tabulation of atomic form factors, photoelectric absorption and scattering cross section, and mass attenuation coefficients in the vicinity of absorption edges in the soft X-Ray (Z=𝑍absentZ=italic_Z =30–36, Z=𝑍absentZ=italic_Z =60–89, E=0.1𝐸0.1E=0.1italic_E = 0.1 keV – 10 keV), addressing convergence issues of earlier work, J. Phys. Chem. Ref. Data 29, 597 (2000).
  • Björck and Andersson [2007] M. Björck and G. Andersson, GenX: an extensible X-ray reflectivity refinement program utilizing differential evolution, J. Appl. Crystallogr. 40, 1174 (2007).
  • Goh et al. [2005] K. E. J. Goh, L. Oberbeck, and M. Y. Simmons, Relevance of phosphorus incorporation and hydrogen removal for Si:P δでるた𝛿\deltaitalic_δでるた-doped layers fabricated using phosphine, Phys. Status Solidi A 202, 1002 (2005).
  • Keizer et al. [2015] J. G. Keizer, S. Koelling, P. M. Koenraad, and M. Y. Simmons, Suppressing segregation in highly phosphorus doped silicon monolayers, ACS Nano 9, 12537 (2015).
  • Flechsig et al. [2010] U. Flechsig et al., Performance measurements at the SLS SIM beamline, AIP Conf. Proc. 1234, 319 (2010).
  • Staub et al. [2008] U. Staub et al., Polarization analysis in soft X-ray diffraction to study magnetic and orbital ordering, J. Synchrotron Radiat. 15, 469 (2008).