Startseite Naturwissenschaften Electrical scanning probe microscopy of electronic and photonic devices: connecting internal mechanisms with external measures
Artikel Open Access

Electrical scanning probe microscopy of electronic and photonic devices: connecting internal mechanisms with external measures

  • Dayan Ban EMAIL logo , Boyu Wen , Rudra Sankar Dhar , Seyed Ghasem Razavipour , Chao Xu , Xueren Wang , Zbig Wasilewski und Sinjin Dixon-Warren
Veröffentlicht/Copyright: 27. Oktober 2015
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

The inner workings of semiconductor electronic and photonic devices, such as dopants, free charge carriers, electric potential, and electric field, are playing a crucial role in the function and performance of the devices. Electrical scanning probe microscopy (SPM) techniques have been developed and deployed to measure, with nanometric spatial resolution and high quantitative accuracy, the two-dimensional profiles of dopant, potential, electric field, and free carrier distribution, within unbiased and/or operating electronic and photonic devices. In this review paper, we summarize our latest SPM experimental results, including the scanning spreading resistance microscopy and scanning capacitance microscopy of terahertz quantum cascade lasers, scanning capacitance microscopy of non-volatile memory devices, scanning voltage microscopy of terahertz quantum cascade lasers, and scanning voltage microscopy of interband cascade lasers. Interpretation of the measured quantities are presented and calibrated, demonstrating that important internal physical quantities and inner mechanisms of device operation can be uncovered. It reveals that the novel SPM techniques would find more applications to the emerging semiconductor quantum devices and nanoelectronics.

1 Introduction

Atomic force microscopy (AFM)-based scanning probe microscopy (SPM) [16] is capable of resolving key inner workings in two dimensions (2D) to delineate quantitatively the transverse cross sections of electronic and photonic devices at nanometer scales. In a conventional atomic force microscope, a sharp probe scans across a surface in physical contact. Feedback control is employed to maintain a constant contact force. A topographic profile of the surface is thereby acquired on the nanometer scale [7, 8].

By employing conductive AFM probes, AFM-based techniques have been extended to probe and map the electronic properties of a semiconductor surface [5, 911]. Two of the most promising techniques, scanning spreading resistance microscopy (SSRM) and scanning capacitance microscopy (SCM), are now commercially available for 2D dopant profiling by the microelectronic and optoelectronic industries. In SSRM, a conductive AFM probe is scanning over a semiconductor device under test in contact mode, and the electric resistance between the probe tip and the device back contact is measured in real time. This resistance is called spreading resistance as it is calculated by dividing the constant tip bias over the resultant spreading current from the probe tip (which is proportional to the local carrier concentration underneath the probe-device contact point). The 2D carrier density profile can therefore be derived from SSRM scans. In SCM, a thin dielectric layer that covers the scan area on the sample surface is typically needed in order to form an effective metal-insulator-semiconductor capacitor between the AFM probe tip and the sample under test. The capacitance-voltage behavior of this effective capacitor is employed to derive the 2D local carrier concentration profile of the sample. Both SSRM and SCM have been applied to III–V compound semiconductor structures and devices [1214]. Nevertheless, these two SPM techniques can only be deployed on unbiased devices because zero device bias is required to prevent multicircuit problems. The techniques therefore delineate only devices at equilibrium.

In order to observe directly the internal behavior of operating quantum optoelectronic devices, a new SPM technique – scanning voltage microscopy (SVM) – has been developed [15]. In SVM, a 2D voltage map is obtained on the cross section of an actively driven semiconductor device. The voltage is typically measured using a very high impedance voltmeter that is connected to a conductive AFM cantilever probe when the probe is scanning over the cross section [16]. Several applications of SVM to diode lasers have been reported, including the direct observation of the internal revelation of current blocking breakdown in a buried heterostructure quantum-well laser [17], the effect of current lateral spreading inside actively biased ridge waveguide lasers [18], and anomalously high series resistance encountered in ridge waveguide lasers [19]. In recent years, new electronic and photonic devices are constantly emerging, such as non-volatile memory (NVM) [20, 21], terahertz quantum cascade lasers (QCLs) [2235] and interband cascade lasers (ICLs) [3641]. The applications of the advanced SPM techniques have naturally been extended to the development of these novel devices in hope of connecting internal mechanisms with external measures [4245].

The article begins by introducing the experimental setups and test devices in the next section. Advances in quantitative dopant profiling of unbiased terahertz QCLs through SSRM and SCM are reviewed in the third section. The fourth section examines the application of SCM to directly read back electrical charges from NVM-based flash devices. The fifth and sixth sections review progress in SVM, in particular, revealing the formation and evolution of electric fields of actively biased terahertz QCLs and quantitatively resolving the dynamic charge carrier distribution in operating ICLs. Finally, the last section summarizes the authors’ findings and the future application of various AFM techniques to nanoscale electronics and photonics.

2 Experimental setups and test devices

The SSRM and SCM scans were performed in a commercial AFM microscope manufactured by Digital Instruments (DI, Dimension 3100, Veeco Instruments, Inc., New York, NY, USA). Boron-doped diamond-coated silicon cantilever tips with a spring constant (42 N/m, Model No. DDESP-10, Bruker AFM Probes, Camarillo, CA, USA) are used in SSRM and SCM measurements in ambient condition. In SSRM mode, a DC bias voltage is applied to the AFM tip and the resulting spreading current through the sample is measured as a function of the tip position on the surface. The spreading current is roughly proportional to the product of the local free carrier concentration and the carrier mobility near the point of probe contact. The SSRM option implements a sensitive six-decade (1 mA–1 nA) logarithmic amplifier to measure the spreading current. The SSRM measurements were performed under contact-mode AFM feedback conditions. The tip was scanned at 0.25–2 μm/s speed, and the sample bias voltage was varied from +1.0 to 3.5 V DC [43].

In SCM mode, the capacitance is measured by using an ultra-high-frequency resonant capacitance sensor, which is connected to the conductive probe tip via a transmission line. Under the direct influence of a sinusoidal bias applied to the sample, charge carriers are periodically depleted or accumulated at the semiconductor surface underneath the AFM probe tip contact point, leading to the change of the capacitance. As the tip scans over the semiconductor surface, the SCM dC/dV signal is measured, which is employed to derive the local free carrier concentration of the sample [44].

The SVM measurements were performed in a commercial cryogenic-temperature atomic force microscope (Attocube) [44, 45]. In SVM mode (Figure 1), a device under test is mounted on a sample holder (heat sink) inside the AFM microscope head. The microscope head is then sealed in a vacuum tube (pumped down to 1×10-4 mbar), which is submerged into a chamber full of liquid nitrogen. A small amount of helium gas is purged into the vacuum tube to an internal pressure of 200 mbar, enabling sufficient heat exchange between the heat sink and the tube wall. The device is then cooled down to 77 K during the SVM scans. A conductive AFM probe is scanned in contact mode over the transverse cross section of the device to pick up the voltage signal [46]. The device is actively biased, and the bottom electrical contact is set as voltage reference. Voltage signals from the AFM probe is directly fed back to and recorded by the AFM data acquisition system through a built-in analog-digital-converter (ADC) channel (voltage resolution, 0.1 mV). The scanning is executed at low frequencies with a slow scan speed of 100–500 nm/s, complying with the synchronization requirement. The applied force to the SVM cantilever probe is in the range of 2–5 μN, depending on measurement requirements. The contact force is set high enough to ensure that the probe is able to penetrate mechanically into the native oxide layer on the semiconductor so as to establish a good electric contact non-destructively, although special care is taken to avoid damaging the laser emission facet. Each 2D SVM image consists of 512 line scans, and each line scan consists of 512 sampling points.

Figure 1: Experimental details. The SVM setup is as follows: the device under test is mounted to a cold finger, and the cold finger is mounted to the piezoelectric stage of the cryogenic atomic force microscope (Attocube, ACS 500). The AFM microscope head is sealed in a vacuum tube (vacuumed and filled with low-pressure helium gas) and submerged into a chamber that is filled with liquid nitrogen. The cleaved laser facet (uncoated) is placed upward, on which the AFM conductive probe scans in contact mode. The device is biased in pulsed mode using an Avtech pulse generator, with a pulse width of 3.5 μm and repetition rate of 100 Hz. The voltage signal picked up by the AFM probe is fed directly into the AFM data acquisition system through a built-in ADC panel 2. The high impedance of the ADC panel minimizes the perturbation to device operation due to the physical contact of the AFM probe across device active region. The noise floor in voltage measurement is ~0.1 mV. In the SVM experiments, the scan range is set to 11×11 μm2 (for rough scans) or 512×512 nm2 (for high-resolution scans). The scan speed is set accordingly to synchronize the SVM voltage reading at each sampling point with the device bias pulses.
Figure 1:

Experimental details. The SVM setup is as follows: the device under test is mounted to a cold finger, and the cold finger is mounted to the piezoelectric stage of the cryogenic atomic force microscope (Attocube, ACS 500). The AFM microscope head is sealed in a vacuum tube (vacuumed and filled with low-pressure helium gas) and submerged into a chamber that is filled with liquid nitrogen. The cleaved laser facet (uncoated) is placed upward, on which the AFM conductive probe scans in contact mode. The device is biased in pulsed mode using an Avtech pulse generator, with a pulse width of 3.5 μm and repetition rate of 100 Hz. The voltage signal picked up by the AFM probe is fed directly into the AFM data acquisition system through a built-in ADC panel 2. The high impedance of the ADC panel minimizes the perturbation to device operation due to the physical contact of the AFM probe across device active region. The noise floor in voltage measurement is ~0.1 mV. In the SVM experiments, the scan range is set to 11×11 μm2 (for rough scans) or 512×512 nm2 (for high-resolution scans). The scan speed is set accordingly to synchronize the SVM voltage reading at each sampling point with the device bias pulses.

The devices under test in the SPM measurements include terahertz QCLs, NVM devices, and ICLs. SSRM, SCM, and SVM measurements were performed on a GaAs-Al0.25Ga0.75As terahertz QCL (V843) with an indirect pumping scheme [29]. The active region of the terahertz QCL consists of 276 cascade modules, each 36.2 nm long. Each module consists of four wells and four barriers with nominal layer thicknesses of 62.5/10.9/66.5/22.8/84.8/9.1/61/44 Å – the barriers are indicated in bold font. The injection barrier (44 Å) is δ-doped with Si to η2D=3.25×1010 cm-2 at the center. The laser devices are processed into metal-metal waveguide ridges 150 μm wide and ~1 mm long. The device emits at 3.2 THz and operates up to a heat-sink temperature of 138 K in pulsed mode. Figure 2 shows a temperature-dependent light-current density (L-J, 10–138 K) and voltage-current density (V-J, 10–250 K) characterizations. All parameters refer to pulsed operation at a repetition rate of 1 kHz and a pulse width of 250 ns for L-J-V measurements.

Figure 2: Light-current density-voltage (L-J-V) characteristics of the terahertz QCL (V843). The J-V curve in (A) is measured at 10 K. At V1=~5.5 V, the J-V curve shows a shoulder, corresponding to the e-1 tunneling resonance. At V2=~9.4 V, the J-V curve shows another shoulder, corresponding to the e-2 tunneling resonance. At V3=~21.8 V, the J-V curve shows negative differential resistance (NDR), corresponding to the misalignment of the e-i tunneling resonance. The device lases up to 138 K in pulsed mode. The pulse width is 250 ns and the pulse repetition rate is 1 kHz in the measurements. No perceptible difference on device performance is observed before and after the SVM scans. (B) J-V curves at temperature range from 10 to 250 K.
Figure 2:

Light-current density-voltage (L-J-V) characteristics of the terahertz QCL (V843). The J-V curve in (A) is measured at 10 K. At V1=~5.5 V, the J-V curve shows a shoulder, corresponding to the e-1 tunneling resonance. At V2=~9.4 V, the J-V curve shows another shoulder, corresponding to the e-2 tunneling resonance. At V3=~21.8 V, the J-V curve shows negative differential resistance (NDR), corresponding to the misalignment of the e-i tunneling resonance. The device lases up to 138 K in pulsed mode. The pulse width is 250 ns and the pulse repetition rate is 1 kHz in the measurements. No perceptible difference on device performance is observed before and after the SVM scans. (B) J-V curves at temperature range from 10 to 250 K.

The core active region of the ICL sample (R090) has eight cascade modules, each consisting of one active section (where the stimulated emission occurs) and one carrier injection section. There are two InAs quantum wells (for electrons) and two GaInSb quantum wells (for holes) in each period of the active section [47]. Micrometer-thick undoped InAs cladding layers sandwich the eight cascade modules from the top and bottom (see Figure 3). The doping concentration in n-doped quantum wells in the electron injection region is 4.3×1018 cm-3. This heavy doping concentration is employed to prevent imbalance of electrons and holes in the active section of the device [40]. The ridge waveguide of the laser (R090) is 10 μm wide and 1 mm long with both facets uncoated. The maximum lasing temperature of the ICL device in pulse mode is 248 K (at which the threshold voltage is 3.8 V).

Figure 3: Schematic diagram of the layer structure of the ICL sample. The interband cascade region consist of eight nominally identical modules; each module contains a 26.9-nm-thick active and hole injection section and a 21.3-nm-thick electron injector section. Each active section consists of two InAs electron wells and two GaInSb hole wells. In order to achieve rebalancing of electrons and holes in the active section, some wells in the InAs-AlSb(As) electron injector section are n+-doped at 4.3×1018 cm-3.
Figure 3:

Schematic diagram of the layer structure of the ICL sample. The interband cascade region consist of eight nominally identical modules; each module contains a 26.9-nm-thick active and hole injection section and a 21.3-nm-thick electron injector section. Each active section consists of two InAs electron wells and two GaInSb hole wells. In order to achieve rebalancing of electrons and holes in the active section, some wells in the InAs-AlSb(As) electron injector section are n+-doped at 4.3×1018 cm-3.

The NVM device employed in the SCM measurements was a Texas Instruments-based microcontroller unit with embedded 512 KB NOR flash. The structure is such that the two arrays of transistors [two floating gates (FGs)] are connected to a common ground so that each array of transistor is connected to the bit line consisting of a series of two FGs. Thus, each cell in the NOR flash has one end connected directly to ground, while the other end is connected directly to a bit line as shown in Figure 4. This is a NOR flash gate; when one of the word lines (connected to the cell’s control gate) is brought high, the corresponding storage transistor acts to pull the output bit line low [48, 49]. The bit line selects the programmed word for each word line and charges the transistor as 1b or 0b depending on the program performed on the device. Each word line transistor is connected to an array on “n” FG transistors (FGTs), as shown in Figure 4, so the first logical bit is simultaneously written in all the FGTs connected to the first word line transistor, the second bit upon the FGTs linked to the second world line transistor, and so on. This ensured that a set of two FGs has the same program data, either “1” or “0”. The NOR flash memory block has a physical measurement of 1.4 mm×2.0 mm in the embedded circuitry of the die from Texas Instruments. The memory area is divided into eight blocks, each having 64 KB of usable memory, and combining to a total 512 KB memory in the microcontroller unit. The two FGs on the transistors have a gate length of 0.22 μm. Figure 5 presents the SEM cross-sectional view of the embedded NOR flash device, showing the structural details of the device, the metal layers, the poly, the tunnel oxides, and the control and FGs.

Figure 4: Schematic of SanDisk based NAND flash memory. (A) Cell structure showing the functionality of transistors in Word data from line 0 to line n, with the Ground Select and Bit Line transistors as the control gate transistors. (B) NMOS transistors representing the cell structure in (A).
Figure 4:

Schematic of SanDisk based NAND flash memory. (A) Cell structure showing the functionality of transistors in Word data from line 0 to line n, with the Ground Select and Bit Line transistors as the control gate transistors. (B) NMOS transistors representing the cell structure in (A).

Figure 5: SEM cross-sectional image of the Texas Instruments-based NOR flash memory showing the transistors. The control and the FGs along with the silicide and metal layers have been marked.
Figure 5:

SEM cross-sectional image of the Texas Instruments-based NOR flash memory showing the transistors. The control and the FGs along with the silicide and metal layers have been marked.

3 SSRM and SCM of terahertz QCLs (V843)

3.1 SSRM

Figure 6 shows a typical 2D SSRM spreading resistance image over an area of 5×12 μm2 on the uncoated facet of the terahertz QCL (V843), which has a metal-semiconductor-metal ridge waveguide structure. In the SSRM scan, the tip bias is 1 V (DC) and the scan rate is 0.5 Hz. The top metal contact layer, the 10-μm-thick quantum-well active region, and the bottom metal contact layer can be delineated in this rough SSRM scan due to their substantial contrast in material electric conductivity. Nevertheless, individual quantum wells (GaAs) and barrier layers (Al0.25Ga0.75As) cannot be resolved because the thickness of each layer (<10 nm) is even smaller than the scan step (12 μm/512=23.4 nm).

Figure 6: The SSRM image of the transverse cross section of the emission facet of the terahertz QCL device (V843). The DC tip bias is 1.0 V, the scan rate is 0.5 Hz, and the scan area is 5×12 μm2.
Figure 6:

The SSRM image of the transverse cross section of the emission facet of the terahertz QCL device (V843). The DC tip bias is 1.0 V, the scan rate is 0.5 Hz, and the scan area is 5×12 μm2.

A high-spatial-resolution SSRM image over an area of 1×0.5 μm2 of the active region of the terahertz QCL is shown in Figure 7. Fourteen periodic thin lines are delineated with an even spacing of 35.7 nm, which is very close to the thickness of one cascade module of the device (the 10 μm active region of the device comprises 276 identical cascade modules, 10 μm/276=36.2 nm). The SSRM signal contrast results from the non-uniform doping profile of the active region – each quantum cascade module is δ-doped with Si dopants. As higher doping concentration yields lower spreading resistance, each thin line in the SSRM image reveals the exact location of δ-dopants in one module and can be used as a “ruler” for measuring the thickness of one cascade module. The small discrepancy (35.7 nm vs. 36.2 nm) may be attributed to experimental calibration error in SSRM. Figure 7 demonstrates that SSRM can be used to delineate individual quantum cascade modules; however, individual quantum wells (GaAs) and barrier layers (AlGaAs) in one module cannot be further resolved. This is because the AFM tip radius (~20 nm) imposes a limited spatial resolution, and GaAs and Al0.25Ga0.75As (small Al composition) have a low material contrast in SSRM.

Figure 7: The zoomed-in SSRM image over the active region of V843. The scan area is 1×0.5 μm2, and the scan rate is 0.25 Hz. This high-solution image delineates 14 quantum cascade modules, with each individual module thickness of ~36 nm.
Figure 7:

The zoomed-in SSRM image over the active region of V843. The scan area is 1×0.5 μm2, and the scan rate is 0.25 Hz. This high-solution image delineates 14 quantum cascade modules, with each individual module thickness of ~36 nm.

The measured spreading resistance at different tip bias (1.0, 1.5, 2.5, and 3.5 V) is plotted as a function of lateral distance from the top metal contact in Figure 8. At higher tip bias, the measured spreading resistance is lower, showing a dependence of the SSRM resistance on the tip bias [10, 12, 16, 50]. The SSRM resistance over the metal layer is the lowest due to its highest electric conductivity. The measured SSRM resistance over the active region varies from 6.1 to 3.2 MΩ for different tip bias voltages because of the low average doping concentration of the active region, which is around 9.0×1015 cm-3. A small negative slope is observed in the SSRM curves over the active region that is in close proximity to the metal layer (see Figure 8). This can be attributed to the effect of nearby metal-semiconductor interface on SSRM current spreading distribution, which has been modeled by De Wolf et al. [51]. Figure 8 shows there are periodic dips (downward peaks) in the SSRM resistance curve every 35.7 nm. As mentioned earlier, those dips (lower SSRM resistance) stem from the high contrast of the δ-doped layer (with a sheet density 3.25×1010 cm-2) comparing to its surrounding undoped layers. Similar signal contrast due to different doping profile has also been reported in SCM scans between n- and p-type materials [52]. Some of the measured SSRM resistance results are summarized in Table 1, which will be used to quantify the average doping concentration of the active region of the device.

Figure 8: 1D SSRM-measured resistance profile near the top metal-semiconductor active region interface of V843 at different DC tip biases of 1.0, 1.5, 2.5, and 3.5 V.
Figure 8:

1D SSRM-measured resistance profile near the top metal-semiconductor active region interface of V843 at different DC tip biases of 1.0, 1.5, 2.5, and 3.5 V.

Table 1

SSRM resistance across the active region of a terahertz QCL device at DC tip bias voltages of 1.0, 1.5, 2.5, and 3.5 V.

DC tip bias1.0 V1.5 V2.5 V3.5 V
n-Type active region6.14×106 Ω5.48×106 Ω3.99×106 Ω3.18×106 Ω

In order to estimate the doping concentration from SSRM measurement, the AFM probe has to be calibrated by using a standard semiconductor sample with staircase doping profile under the same condition [10]. Figure 9 shows four calibration curves, showing the measured SSRM resistance as a function of doping concentration at DC tip bias voltages of 1.0, 1.5, 2.5, and 3.5 V, respectively. As expected, the SSRM resistance monotonically decreases as the doping concentration increases over a range of 1015–1019 cm-3. The calibration curves will be used shortly to derive the average doping concentration of the QCL active region.

Figure 9: Four SSRM calibration curves obtained at different DC tip biases (1.0, 1.5, 2.5, and 3.5 V), correlating the SSRM measured resistance to the doping concentration in a GaAs standard sample with staircase doping profile. The doping concentration of the GaAs standard sample is obtained from second ion mass spectroscopy. The error bars are ±10% and ±5% for x- and y-axis, respectively.
Figure 9:

Four SSRM calibration curves obtained at different DC tip biases (1.0, 1.5, 2.5, and 3.5 V), correlating the SSRM measured resistance to the doping concentration in a GaAs standard sample with staircase doping profile. The doping concentration of the GaAs standard sample is obtained from second ion mass spectroscopy. The error bars are ±10% and ±5% for x- and y-axis, respectively.

3.2 SCM

The terahertz QCL device (V843) is also examined by using SCM, which introduces less physical damages to the sample surface due to weak force being applied by the AFM probe during the scans. A typical SCM image from V843 is presented in Figure 10. During the open-loop mode scans, the AC bias voltage is 1.0 V and the scan rate is 0.5 Hz. The SCM image can also resolve the metal layers and the 10-μm-thick semiconductor active region, but with a lower spatial resolution comparing to the SSRM image. This lower spatial resolution is attributed to the fact that the lateral distribution of the depletion region in semiconductor underneath the probe contact point introduces a substantial shadow effect to smear the SCM image [12].

Figure 10: SCM image of the transverse cross section of the emission facet of V843. The AC tip bias is 1.0 V, the scan rate is 0.5 Hz, and the scan area is 5×12 μm2.
Figure 10:

SCM image of the transverse cross section of the emission facet of V843. The AC tip bias is 1.0 V, the scan rate is 0.5 Hz, and the scan area is 5×12 μm2.

Figure 11 shows a zoomed-in SCM image over an area of 1×0.5 μm2 of the active region with a scan rate of 0.25 Hz and an AC bias of 1.0 V under an open-loop amplitude mode. The δ-doped layers can still be resolved as 14 thin lines in the SCM image, but appear not as sharply as the ones resolved in the SSRM image due to the poorer spatial resolution of SCM. The spacing between the 14 thin lines is measured to be 35.7 nm, which is the same as the one obtained from SSRM scans.

Figure 11: The zoomed-in SCM image over the active region of V843. The scan area is 1×0.5 μm2 and the scan rate is 0.25 Hz. Fourteen quantum cascade modules are delineated.
Figure 11:

The zoomed-in SCM image over the active region of V843. The scan area is 1×0.5 μm2 and the scan rate is 0.25 Hz. Fourteen quantum cascade modules are delineated.

Even though efforts have been made to attain a direct correlation between the SCM signals and semiconductor doping concentration [5355], there are no well-established and reliable models or algorithm for directly converting the measured SCM signals to the doping concentration. A standard semiconductor sample with staircase doping profile is used to calibrate the SCM approach, as was done in SSRM. Figure 12 presents the five calibration curves for SCM from the n-type GaAs staircase doping profile sample. Contrary to the SSRM calibration curves, the measured SCM signal (dC/dV) monotonically increases with the increase of the doping concentration over the range of 1×1015–4×1018 cm-3. The SCM calibration curves, together with the measured SCM data from the terahertz QCL, will be used to derive the average doping concentration in the active region of the V843 device.

Figure 12: Five SCM calibration curves obtained at different AC tip biases (0.5, 0.75, 1.0, 1.25, and 2.0 V), correlating the SCM signal to the doping concentration in the GaAs standard sample with staircase doping profile.
Figure 12:

Five SCM calibration curves obtained at different AC tip biases (0.5, 0.75, 1.0, 1.25, and 2.0 V), correlating the SCM signal to the doping concentration in the GaAs standard sample with staircase doping profile.

3.3 Calibration results

Combining the SSRM calibration curves, the SCM calibration curves, and the measured data from the terahertz QCL device, the average doping concentration of the 10-μm-thick active region of V843 are derived and summarized in Table 2. Shown together is the nominal doping concentration of the device. The derived doping concentration exhibits a reasonably good agreement with the design value (9.0×1015 cm-3), while the SSRM-based values show a larger discrepancy (within 25%) and the SCM-based values are more consistent (within ~10%). The discrepancy between the measured doping concentration and the design value could mainly be attributed to system errors in the scans, the non-linear behavior of the tip-sample Schottky contact, and the non-identical surface condition between the terahertz QCL samples and the calibration GaAs staircase samples [10, 12, 56]. The experimental results demonstrate that SSRM and SCM can delineate 2D semiconductor structures with a high spatial resolution and resolve the doping profiles in semiconductor layers with a certain quantitative accuracy.

Table 2

Comparison of average doping concentrations derived from calibrated SSRM and SCM measurements, and the average designed doping concentration values of the active region of the terahertz QCL device for different DC and AC bias voltages, respectively.

Terahertz QCL structureNominal average doping conc. (×1015)SSRM DC=1.0 V (×1015)SSRM DC=1.5 V (×1015)SSRM DC=2.5 V (×1015)SSRM DC=3.5 V (×1015)SCM AC=0.5 V (×1015)SCM AC=0.75 V (×1015)SCM AC=1.0 V (×1015)SCM AC=1.25 V (×1015)SCM AC=2.0 V (×1015)
n-Type active region9.00 (±10%)8.00 (±10%)6.80 (±10%)6.00 (±10%)6.80 (±10%)8.10 (±10%)9.70 (±10%)10.20 (±10%)10.00 (±10%)9.50 (±10%)

4 SCM of NVM devices

4.1 Programing the NVM devices and sample preparation

To prepare the SCM experiments, two 512 KB Texas Instrument NVMs (NOR flash devices) are selected and programmed with the word data “AA”=“10101010” and “CC”=“11001100”, respectively, all over their memory blocks. The static charges are stored at the FG-gate oxide interface in the NVM devices, an approach that modifies the threshold voltage and the channel carrier distribution of the FGT [57]. In the FGT channel, the electron density stored in its FG by capacitance coupling through the gate oxide is actually the coupled image of the hole density [58, 59]. In order to probe the carrier density located on the FGT channels by using SCM, the conductive AFM probe has to scan close enough (<300 nm) to the transistor channels but without discharging the FGs. The samples under test therefore need to be deprocessed in a very definite way. A backside approach is adopted for the sample preparation [57], in which the tunnel oxide is exposed from the backside by deprocessing the sample. Figure 13 schematically shows the main steps of this backside approach. The Si substrate is polished to reduce its thickness from around 15 μm down to 0.5 μm, and then a final polishing is performed to achieve the desired silicon thickness of 30–200 nm. A thinner silicon substrate would yield better SCM scan results, but it may also lead to charge loss due to charge tunneling. The FG potentials can then be measured through this tunnel oxide [58]. It is practically impossible to achieve a uniformly thin silicon layer (with a thickness <200 nm) all over the memory device through polishing. A practically feasible approach is to adopt a bevel configuration with a small slope (~0.15%), where only a small area (≈80 μm2 large) is usable for the SCM measurements, as marked by a white box and shown in Figure 14. A diamond-coated AFM probe tip is employed in the SCM scans and both DC and AC voltages are applied to the AFM probe in the contact mode [48, 5861].

Figure 13: Schematic diagram of the sample preparation process performed on the NVM devices. (A) Sample is glued to the AFM sample holder using silver epoxy, making it ready for backside deprocessing. (B) Rough polishing on the backside of sample to remove the bulk silicon substrate until a thickness of 0.5 μm. (C) Fine polishing to slowly etch the silicon substrate down to a thickness of 30–200 nm.
Figure 13:

Schematic diagram of the sample preparation process performed on the NVM devices. (A) Sample is glued to the AFM sample holder using silver epoxy, making it ready for backside deprocessing. (B) Rough polishing on the backside of sample to remove the bulk silicon substrate until a thickness of 0.5 μm. (C) Fine polishing to slowly etch the silicon substrate down to a thickness of 30–200 nm.

Figure 14: An optical microscope image of the final prepared sample, which is ready for SCM scanning on the backside, showing the thin silicon layer in a bevel pattern.
Figure 14:

An optical microscope image of the final prepared sample, which is ready for SCM scanning on the backside, showing the thin silicon layer in a bevel pattern.

4.2 Stored charge carriers

The buried charge carriers stored in the FGT channels of the embedded NOR flash memory device can now be detected in SCM scans. The capacitance resonator of the SCM has a high sensitivity of 10-22 F/√Hz to carriers, allowing probing electric charges as small as a single electron [62]. The SCM signals are measured with 2 V AC and -1.8 V DC bias applied to the AFM tip. Figure 15 shows such SCM images over the deprocessed NOR flash devices. The big yellow-colored regions represent the common sources of the FGs, while the small yellow spots (underneath the source regions, with fading color) are individual drains. The drains appear much dimmer in the SCM images comparing to the source regions because the drains are further away from the AFM probe tip. When the FGs are charged, the drains become invisible in the SCM images because the signal from the stored charge overshadows that of the drains. The gate length is measured to be 0.22 μm from the SCM images.

Figure 15: Two SCM scan images over a 2.50 μm distance of the embedded NOR flash memory device. The preprogrammed word data (A) “AA”=“1010” in binary form, and (B) “CC”=“1001” in binary form, can be read directly from the SCM images. The stored dynamic charges are identified to be present on FGs for “1b” while absent on the FGs for “0b”.
Figure 15:

Two SCM scan images over a 2.50 μm distance of the embedded NOR flash memory device. The preprogrammed word data (A) “AA”=“1010” in binary form, and (B) “CC”=“1001” in binary form, can be read directly from the SCM images. The stored dynamic charges are identified to be present on FGs for “1b” while absent on the FGs for “0b”.

In Figure 15, bright blue-colored regions are observed, which signify the stored electric charges on the FGs and represent a “1b” or ON state. The FGs without stored charges appear in pink color, representing a “0b” or OFF state. The ON and OFF states in Figure 15A and B are marked with black and white circles, respectively. In Figure 15A, binary bits of “1010” can be observed in a repeatable pattern throughout the FGTs, corresponding to the NVM devices that are preprogrammed with the word data “AA”. In Figure 15B, binary bits of “1001” can be observed, corresponding to the NVM devices that are preprogrammed with the word data “CC”. The results demonstrate that SCM is capable of reading the stored charge carriers with submicrometer spatial resolution and high accuracy.

It can also be observed that the appearance of the FGs at the ON state is not exactly identical in the SCM images. As the brightness of the blue-colored region quantitatively reflects the density of the charge carriers stored in FGs, the SCM images provide a direct visualization of the 2D stored charge density profile (Figure 15), showing the charge carrier density may vary substantially across the FGs at the ON state, which should nominally be identical to one another. This variation is revealed more clearly in the cross-section analysis, which is shown in Figure 16. The one-dimension dC/dV signal in Figure 16 is obtained by averaging the line scans of the region marked with white dotted lines in Figure 15B. Over the distance range of 0.5–2.0 μm, there are two positive peaks and two negative peaks. The positive peaks correspond to “1b”, indicating accumulation of charge carriers. The negative peaks correspond to “0b”, representing no charge carriers. The data is therefore read as “1001”, correlating well to the preprogrammed word of “CC” in the NOR flash device. As mentioned earlier, the variation in the measured dC/dV signal of the ON states is attributed to the difference in the quantity of the charge carriers stored in the FGs. The signal-to-noise ratio in Figure 16 is estimated to be ~11, sufficiently high for reliably determining the digital bits stored in the FGs.

Figure 16: SCM cross-section analysis of Figure 15B for the 2.5 μm distance. The positive peaks represent FGs with stored charges, while the negative peaks represent FGs without.
Figure 16:

SCM cross-section analysis of Figure 15B for the 2.5 μm distance. The positive peaks represent FGs with stored charges, while the negative peaks represent FGs without.

4.3 Repeatability

The stored data in a Texas Instruments 512 KB embedded NOR flash memory devices of 0.22 μm gate length FGTs are successfully read back using the SCM technique. The results show that individual digital bits of “1b” and “0b” can be determined reliably from SCM scans; moreover, SCM images also reveal the different charge carrier densities stored in FGs. The process used for the sample preparation and the SCM reading of the stored data appears to be data non-destructive and repeatable. The DC and AC voltages applied to the tip for charge detection are maintained at a low range ~(-3.0 to -0.1 V) and ~(+1.0 to +3 V), respectively, levels far below from program-erasable high voltages ~(+8 V, +20 V) for the Flash devices. SCM scanning (reading) of sample surfaces are performed for ~4 h continuously in the same area, and no noticeable reduction of charge level is detected [57]. The device is kept in a dark box after the sample preparation in order to retain the programmed charges and minimized charge loss due to external optical illumination [48, 58]. After a long storage time, the SCM readings of the digital bits are still correct and agree well with the programmed data. This demonstrates that the SCM-based read-back process is reliable and data non-destructive.

5 SVM of terahertz QCLs

5.1 Formation and evolution of electric field domain

In a terahertz QCL, each individual quantum cascade module needs to be biased exactly at designed electric field in order to achieve population inversion – a prerequisite for lasing. In many cases, no lasing can be observed because the quantum cascade modules are biased at wrong electric fields potentially due to the formation of electric field domains (EFDs) in the active region – a hypothesis that has long been suspected and could only be verified indirectly from the measurements of active-region photoluminescence spectra or the observation of sawtooth-like current-voltage (I-V) or light-voltage (L-V) curves [6366]. SVM has been shown to resolve individual semiconductor quantum wells and quantitatively measure voltage distribution across semiconductor layers [17, 19], and is thus used to experimentally explore the formation and evolution of EFDs in operating terahertz QCLs.

An indirect pumping-based terahertz QCL laser with uncoated cleaved facets and a metal-metal waveguide [29] is used in the SVM scans. Figure 17A shows a typical 2D voltage profile across the device active region. The device is biased at 12 V in pulsed mode at 77 K. As expected, the measured voltage profile discloses a monotonic drop in voltage from the top metals (at 12 V), through the intermediate layers, to the grounded bottom metals (at 0 V). There is a sharp voltage drop of ~0.8 V at the Schottky junction between the top metal and the top n+ GaAs contact layer. The SVM image also directly visualizes the formation of two EFDs at this bias – one higher EFD close to the top metal layer (greater slope) and one lower EFD close to the bottom metal layer (smaller slope). The inset is the topology of the scanned cross section, showing that the cleaved emission facet is almost atomically flat.

Figure 17: SVM results of an operating terahertz QCL (V843), showing the formation and evolution of EFDs in the active region. (A) An SVM-measured 2D voltage profile across V843 at a forward bias of 12 V (pulsed mode, 3.5 μs pulse width, and 100 Hz repetition rate) at 77 K. Two EFDs (F1 and F2) across the ~10-μm-thick MQW active region are observed. The inset of the figure shows a 2D topology image over the same area. (B and C) 1D voltage profile across the device active region at different biases (2–25 V). (D) Experimental and simulated electrical and optical characteristics of V843 at 77 K and other temperatures (10, 50, 100, and 130 K). (E) The SVM-measured electric field across individual cascade modules in the active region of operating V843 device as a function of applied device bias. In the bias ranges of 10–17 V and 22–25 V, the active region splits to two EFDs and the electric field in each EFD remains constant. Shown together is the partition number of the cascade modules in each EFD measured from SVM. The number of cascade modules in the higher EFD increases, while the number of cascade modules in the lower EFD decreases with the device bias.
Figure 17:

SVM results of an operating terahertz QCL (V843), showing the formation and evolution of EFDs in the active region. (A) An SVM-measured 2D voltage profile across V843 at a forward bias of 12 V (pulsed mode, 3.5 μs pulse width, and 100 Hz repetition rate) at 77 K. Two EFDs (F1 and F2) across the ~10-μm-thick MQW active region are observed. The inset of the figure shows a 2D topology image over the same area. (B and C) 1D voltage profile across the device active region at different biases (2–25 V). (D) Experimental and simulated electrical and optical characteristics of V843 at 77 K and other temperatures (10, 50, 100, and 130 K). (E) The SVM-measured electric field across individual cascade modules in the active region of operating V843 device as a function of applied device bias. In the bias ranges of 10–17 V and 22–25 V, the active region splits to two EFDs and the electric field in each EFD remains constant. Shown together is the partition number of the cascade modules in each EFD measured from SVM. The number of cascade modules in the higher EFD increases, while the number of cascade modules in the lower EFD decreases with the device bias.

One-dimensional SVM-measured voltage profile curves are also measured at different device biases from 2 to 25 V, and the results are presented in Figure 17B and C. No multiple EFDs are observed at low biases (2–9 V). Beyond 10 V, a higher EFD (denoted by dashed lines) emerges from the active region close to the top metal layer and expands linearly with the increase of the device bias to the bottom metal layer at the expense of the lower EFD (solid lines), until it expand over the whole active region (10 μm thick). At the biases between 18 and 21 V, only one slope is observed. Multiple EFDs appear again in the device bias range of 22–25 V, and similar EFD evolution behaviors are observed. The SVM results thus directly reveal the formation and evolution of the EFDs in an operating terahertz QCL. In addition, the electric field (F) of the observed EFDs are quantitatively measured from the slopes [F=average of (ΔVd)] of the voltage profile curves in Figure 17B and C. Table 3 summarizes the measured electric field as a function device bias. Figure 17D shows the experimental and theoretical light-current density-voltage (L-J-V) characteristics of the device under test. It has been found that the electric field of the EFDs measured from SVM correlates very well to those of the important features observed in the L-J-V curves in Figure 17D, which can be well explained by the alignment and misalignment of the discrete energy levels in the quantum wells of the device active region as well as the competition of various current-carrying channels under different device biases [44].

Table 3

SVM-measured electric field in each observed EFD.

Applied device bias (V)F1 (kV/cm)F2 (kV/cm)F3 (kV/cm)F4 (kV/cm)
10.08.58316.767
11.08.57416.767
12.08.57916.767
13.08.57416.763
14.08.58616.763
15.08.58016.770
16.08.57916.770
17.08.58316.767
22.020.95224.347
22.520.95724.347
23.020.96124.347
23.520.95524.347
24.020.95924.347
25.020.95724.347
Average (Eavg)8.5816.7720.9624.35

The electric field values are obtained by linearly fitting the different sections of the voltage curves. The small variation is attributed to small system errors.

The number (nk) of the cascade modules in each EFD section can be calculated by using nk=lk/d (k=1, 2), where lk is the length of each EFD section measured directly from Figure 17B and C, and d is the thickness of one cascade module (d=36.2 nm) [29]. Figure 17E presents the calculated numbers at different device biases, clearly confirming that the boundary of the EFDs linearly shifts from the top to the bottom contact layer as device bias increases over two ranges of 10–17 V and 22–25 V [44].

5.2 High-resolution SVM

By reducing the scan range, the spatial resolution of the SVM scans can be enhanced so that individual quantum cascade modules can be delineated. Figure 18A shows a typical 1D rough SVM scan across the ~10 μm multiple quantum well (MQW) active region at a device bias of 15 V. Figure 18B–D shows high-resolution voltage profiles by zooming in the SVM scans in three 512×512 nm2 regions – one close to the top metal-semiconductor interface, one close to the EFD boundary, and one close to the bottom metal-semiconductor interface. A small voltage dip is observed every ~36 nm in the high-resolution SVM curves. The small voltage dip is attributed to the δ-doping profile (η2D=3.25×1010 cm-2) in the injection barrier, which is the boundary layer between two neighboring modules. As a result, the voltage dips can be used as a “ruler” in SVM scans.

Figure 18: (A) The 1D SVM voltage curve cross the active region of the V843 device (device bias: 15 V, temperature: 77 K). The inset shows three zoomed-in SVM scans spanning 512 nm each. (B–D) are the corresponding further zoomed-in curves in which the small voltage dips at the δ-doped injection barriers can be clearly observed. The bottom curve in each figure is the first-order derivative (|dV/dx|) of the corresponding voltage profile curve.
Figure 18:

(A) The 1D SVM voltage curve cross the active region of the V843 device (device bias: 15 V, temperature: 77 K). The inset shows three zoomed-in SVM scans spanning 512 nm each. (B–D) are the corresponding further zoomed-in curves in which the small voltage dips at the δ-doped injection barriers can be clearly observed. The bottom curve in each figure is the first-order derivative (|dV/dx|) of the corresponding voltage profile curve.

Figure 19 shows high-resolution SVM scans [(A) for 2D and (B) for 1D], which zoom in the EFD boundary region. All scans exhibit two distinct voltage slopes, and the small voltage dips denote individual cascade modules. The periodic spacing measured from these high-resolution SVM scans is ~36.1±0.1 nm, agreeing well with the design value of 36.2 nm. It is also revealed that the EFD boundary (at which the two voltage slopes cross) is located at ~12±0.5 nm away from the small voltage dip (the injection barrier) toward the upstream direction of the electron flux (Figure 19B). Figure 19C shows a series of high-resolution SVM scans by changing the device bias from 14.991 to 15.039 V with an incremental step of ~0.01 V. The experimental results show that the EFD boundary does not shift continuously with the gradual increase of the applied device bias, but rather hops discretely. Only when the device bias is increased by ~30 mV or more does the EFD boundary jump by ~36 nm, which is one cascade module thickness. The minimum bias increase that is needed to flip one complete cascade module from the lower EFD (F=8.58 kV/cm) to the higher EFD (16.77 kV/cm) is 30 mV. This one-module-at-a-time progression of the EFD boundary is therefore convincingly confirmed to be the nanoscopic origin of the reported sawtooth-like current-voltage (I-V) characteristics exhibited by QCLs [67].

Figure 19: High-resolution SVM scans near the boundary of EFDs. (A) The 2D SVM voltage image over a 512×512 nm2 scan area at 15 V and 77 K. (B) A 1D zoomed-in voltage curve over three modules in a row, showing the exact location of the EFD boundary (the turning point of the electric field). (C) 1D SVM voltage profile curves at a series of device biases from 14.991 to 15.039 V, showing the hopping behavior of the EFD boundary. All curves (except the one at 15.001 V) are accumulatively shifted by 0.01 V vertically for clarity.
Figure 19:

High-resolution SVM scans near the boundary of EFDs. (A) The 2D SVM voltage image over a 512×512 nm2 scan area at 15 V and 77 K. (B) A 1D zoomed-in voltage curve over three modules in a row, showing the exact location of the EFD boundary (the turning point of the electric field). (C) 1D SVM voltage profile curves at a series of device biases from 14.991 to 15.039 V, showing the hopping behavior of the EFD boundary. All curves (except the one at 15.001 V) are accumulatively shifted by 0.01 V vertically for clarity.

5.3 Simulation results

In order to understand why the EFD boundary is ~12 nm after the injection barrier, the potential profile, electric field, and charge carrier distribution are calculated by self-consistently solving coupled Schrödinger-Poisson equations. The simulation results across the quantum cascade modules in close proximity of the EFD boundary are presented in Figure 20. The input parameters in the simulation include the lower electric field (F1), the higher electric field (F2), and the voltage drop across seven modules around the EFD boundary, which are derived from a high-resolution SVM scan at 12 V. The calculated band diagram, the potential profile, and the wave functions of the confined energy states across the modules are presented in the lower part of Figure 20, which shows a clear turning point between the higher EFD and lower EFD. The turning point (the EFD boundary) is located in the widest well of the module, coinciding with the lobe of the wave function of the extraction state, which is ~12.3 nm away from the center of the injection barrier layer, confirming the experimental results from the SVM measurements. The widest quantum well accommodates the electron accumulation that is resulted from the discontinuity of the electric field across the EFD boundary [44].

Figure 20: Simulated charge carrier density (ρ/q, top section); electric field (F, middle section); and band diagram, electrical potential, and wavefunctions (bottom section) over quantum cascade modules near the EFD boundary. The simulation results confirm that the EFD boundary is ~12.3 nm away from the δ-doped injection barrier in the transitional module. The device bias used in the simulation is 12 V.
Figure 20:

Simulated charge carrier density (ρ/q, top section); electric field (F, middle section); and band diagram, electrical potential, and wavefunctions (bottom section) over quantum cascade modules near the EFD boundary. The simulation results confirm that the EFD boundary is ~12.3 nm away from the δ-doped injection barrier in the transitional module. The device bias used in the simulation is 12 V.

6 SVM of ICLs

6.1 Voltage profile inside operating ICLs

It has been postulated that the electron density and hole density are not balanced in the active region in an operating ICL, which substantially degrades device performance. By heavily n-doping the electron injection layers, the performance of ICLs has been significantly improved, which was attributed to the rebalancing of electrons and holes [39]. In this section, we present direct experimental evidences obtained from SVM measurements, confirming the rebalancing of electrons and holes in an actively biased ICL with a heavy n-doping profile. The experimental results also show that the charge carrier density in a lasing ICL at low temperatures is pinned above the threshold.

Figure 21A is a 2D voltage profile over a transverse cross-section area (3×3 μm2) on the front emission facet of the ICL (R090) under a forward bias of 3 V. The top metal layer, the top separate confinement layer (SCL, undoped InAs), the interband cascade region (eight cascade modules), and the top SCL layer can be resolved in Figure 21A due to their contrast in electric potential. Most of the voltage drop is observed over the interband cascade region (386 nm in thickness). Figure 21B presents 1D voltage profile curves at different device biases from 0.5 to 5.0 V. The slope of the voltage profile is uniform over the interband cascade region, and it increases proportionally with the increase of device bias. A slight voltage drop is observed at the top contact layer-top SCL layer interface, which is under a local reverse bias even though the device is globally forward biased.

Figure 21: (A) An SVM 2D voltage profile image of an ICL (device bias: 3.0 V, temperature: 77 K). (B) 1D section analysis of the SVM voltage profiles at different device biases from 0.5 to 5.0 V. It shows that the voltage drop mainly occurs across the interband cascade active region of the device.
Figure 21:

(A) An SVM 2D voltage profile image of an ICL (device bias: 3.0 V, temperature: 77 K). (B) 1D section analysis of the SVM voltage profiles at different device biases from 0.5 to 5.0 V. It shows that the voltage drop mainly occurs across the interband cascade active region of the device.

6.2 Electric field profile and carrier density profile

High-resolution SVM scans are achieved by reducing the scan range to delineate individual cascade modules. Figure 22A shows such a high-resolution voltage image over a small area (512×512 nm2), in which eight pairs of voltage dips (denoted by dashed lines) appear with a spacing of ~48 nm over the 386-nm-thick interband cascade region, corresponding well to the eight cascade modules in R090. Calculating the first-order derivative of measured voltage profile against distance (F=-dV/dx) yields the internal electric field (F) inside the MQW active region, which is plotted in Figure 22B. The average electric field in the MQW active region is F=8.64 V/μm (86.4 kV/cm) at a device bias of 3.5 V. Two peaks and two valleys are observed in the electric field curve over each of the voltage dip pair regions. This non-uniform electric field profile is attributed to spatial segregation of electrons and holes, which is expected due to the unique type-II quantum wells of this ICL device. The identical electric field distribution profile among the eight cascade modules (Figure 22B) reflects uniform carrier injection and distribution in each of the modules, ensured by the series connection of the modules in the structure [38, 68]. The measured voltage dip pairs and the derived electric field peaks and valleys are more perceivable in a zoomed-in plot (see Figure 22C). By comparing to the band diagram of the ICL (Figure 22D), it is found that the voltage dips exactly overlap with the InAs-GaInSb MQW active section of the corresponding modules.

Figure 22: (A) A zoomed-in 2D voltage profile across the interband cascade region (scan area: 512×512 nm, temperature: 77 K, and applied bias: 3.0 V). (B) A 1D voltage profile over a distance of 512 nm, together with corresponding electric field (-dV/dx) distribution across the region. (C) A zoomed-in voltage profile over a distance of 63 nm (temperature: 77 K, device bias: 3.5 V), showing two pairs of voltage dips every ~48 nm. The bottom curve shows the corresponding electric field curve. (D) The zoomed-in profile (-d2V/dx2) at 3.5 V, numerically derived from the electric field curve (C). The top curve is the zero-bias band diagram of the MQWs in the active region.
Figure 22:

(A) A zoomed-in 2D voltage profile across the interband cascade region (scan area: 512×512 nm, temperature: 77 K, and applied bias: 3.0 V). (B) A 1D voltage profile over a distance of 512 nm, together with corresponding electric field (-dV/dx) distribution across the region. (C) A zoomed-in voltage profile over a distance of 63 nm (temperature: 77 K, device bias: 3.5 V), showing two pairs of voltage dips every ~48 nm. The bottom curve shows the corresponding electric field curve. (D) The zoomed-in profile (-d2V/dx2) at 3.5 V, numerically derived from the electric field curve (C). The top curve is the zero-bias band diagram of the MQWs in the active region.

Calculating the second-order derivative of V(x) yields the net charge carrier density [n(x)],

(1)n(x)=-ε0εrqd2V(x)dx2, (1)

where ε0 is the vacuum permittivity, εr the relative permittivity, and q the electron charge. Figure 22D presents the calculated -d2V/dx2 results, as well as the band diagram of the active and injection sections at zero bias. It shows two negative peaks and three positives across the InAs-GaInSb active section, which correspond to net negative charge accumulation in the two InAs conductive-band quantum wells and net positive charge in the two GaInSb and the neighboring AlSb valence-band quantum wells. The net charge density is zero in other region. The experimental results based on SVM measurements therefore provide the first-ever direct observation of the electron and hole spatial segregation at nanometer scales in an operating ICL [45].

6.3 Bias dependence and temperature dependence of charge carrier density

By integrating n(x),

(2)σi=in(x)dx, (2)

one can quantitatively calculate the net charge carrier density (σi, i=1, 2, 3, 4, 5) in individual quantum wells (as labeled in Figure 22D). Figure 23A presents the calculated carrier sheet density as a function of the applied bias at 77 K, which shows a quick increase prior to the lasing threshold voltage (3.4 V) and saturation at and above the threshold. The saturation of charge carrier density above the threshold voltage is because the additional injected carriers are completely consumed by stimulated emission [69]. Similar SVM measurements and data analysis are repeated on the same device at room temperature, the results are presented in Figure 23B. The carrier density exhibits monotonic increase over the device bias range from 0 to 4 V. This non-clamping behavior is attributed to the absence of stimulated emission at this temperature (the device’s maximum lasing temperature is 248 K). Table 4 summarizes the derived electric field and sheet carrier density at different device biases. The total hole density (σh) and the total electron density (σe) are in reasonable agreement with each other (~10%, within experimental uncertainty), confirming the maintaining of charge neutrality.

Figure 23: (A) The experimental sheet charge carrier densities at different device biases at 77 K. The threshold voltage at 77 K is 3.4 V. (B) The experimental sheet charge carrier densities at different device biases at room temperature. The device does not lase at room temperature. (C) The experimental total electron density (σ2+σ4) as a function of the internal average electric field. A linear fitting curve (solid line) based on Eq. (3) is plotted for comparison.
Figure 23:

(A) The experimental sheet charge carrier densities at different device biases at 77 K. The threshold voltage at 77 K is 3.4 V. (B) The experimental sheet charge carrier densities at different device biases at room temperature. The device does not lase at room temperature. (C) The experimental total electron density (σ2+σ4) as a function of the internal average electric field. A linear fitting curve (solid line) based on Eq. (3) is plotted for comparison.

Table 4

The measured internal average electric field (F), the experimental charge carrier sheet density (σi, i=1, 2, 3, 4, 5), the total hole sheet density (σh=σ1+σ3+σ5), the total electron sheet density (σe=σ2+σ4), and the theoretical modeling sheet carrier density (σm) by using Eq. (5).

Device bias (V)F (×106 V/m)σ1 (cm-2)σ2 (cm-2)σ3 (cm-2)σ4 (cm-2)σ5 (cm-2)σh (cm-2)σe (cm-2)σm (cm-2)
001.18E+093.43E+092.8E+093.63E+091.18E+095.16E+097.07E+09
0.51.1504.43E+091.29E+101.59E+101.37E+104.43E+092.48E+102.66E+10
1.02.3991.18E+103.23E+104.2E+103.84E+101.18E+106.56E+107.07E+105.54E+10
1.53.6312.95E+108.88E+101.11E+119.29E+102.95E+101.7E+111.82E+111.95E+11
2.04.8695.31E+101.57E+111.9E+111.62E+115.31E+102.96E+113.19E+113.35E+11
2.56.1177.97E+102.22E+112.83E+112.54E+117.97E+104.43E+114.76E+114.76E+11
3.07.3721.15E+112.99E+113.54E+113.39E+111.15E+115.84E+116.38E+116.18E+11
3.38.1331.27E+113.23E+113.58E+113.63E+111.27E+116.12E+116.86E+117.04E+11
3.58.6381.33E+113.31E+113.72E+113.71E+111.33E+116.37E+117.03E+117.56E+11
4.09.9241.39E+113.47E+113.89E+113.88E+111.39E+116.67E+117.35E+118.92E+11
4.511.211.42E+113.55E+113.98E+113.96E+111.42E+116.81E+117.51E+111.03E+12
5.012.501.42E+113.55E+113.98E+113.96E+111.42E+116.81E+117.51E+111.16E+12

mr=0.0355me is used in the calculation, where me is the free electron mass. The device is forward biased at 77 K with a threshold voltage of 3.4 V. The uncertainty of the experimental charge carrier sheet density is estimated to be ±10%. The modeling results are in good agreement with the experimental data (data in boldface).

The quasi-equilibrium charge carrier sheet density has been found to be linearly dependent on electric field (F) under certain conditions [68]:

(3)σh=σemrESMπ2=mrqF×Δd-Eiπ2, (3)

where mr is the reduced effective mass, the reduced Planck constant, and ESM the overlap energy. Δd is the center-of-mass distance between the electron and hole wave function distributions across the type II interface. Figure 23C plots the experimental electron density at 77 K as a function of the internal electric field (F), together with a linear fitting curve based on Eq. (3), showing a very good agreement over the electric field range of 2.399×106 V/m (23.99 kV/cm) to 8.133×106 V/m (81.33 kV/cm) (corresponding to applied device biases from 1.0 to 3.3 V). Two fitting parameters are obtained as Δd=7.6 nm and Ei =14.6 meV.

The temperature dependence of the charge carrier sheet density is also investigated by conducting SVM measurements on R090 at a constant bias (3.8) for different temperatures followed by corresponding data analysis. The results are presented in Figure 24. The threshold current increases slowly with temperature from 60 to 170 K, and then rapidly increases till the maximum lasing temperature. The measured total electron sheet density and hole density shows a linear dependence on temperature from 77 to 240 K (increases by only ~15%). At temperatures above the maximum lasing temperature, the total carrier density increases rapidly by ~140% from 240 to 298 K. This fast increase observed in carrier density above maximum lasing temperature is attributed to the absence of stimulated emission. The SVM results have shown that the optical gain (proportional to carrier density) has only modest increase from 77 to 240 K. Nevertheless, the threshold current increases rapidly in that temperature range with a characteristics temperature T0=~29 K, which is likely attributed to a fast reduction in carrier lifetime due to stronger non-radiative Auger recombination process at elevated temperatures [70, 71].

Figure 24: The active-region charge carrier sheet densities (σh and σe) measured at 3.8 V from SVM as a function of temperature; shown together is the experimental threshold current as a function of temperature.
Figure 24:

The active-region charge carrier sheet densities (σh and σe) measured at 3.8 V from SVM as a function of temperature; shown together is the experimental threshold current as a function of temperature.

7 Conclusions

By applying SPM techniques to a few emerging advanced electronic and photonic devices, the inner workings such as doping profiles, electric potential profile, electric field profile, and charge carrier profile in biased or unbiased devices are measured non-invasively at nanometer scales. The experimental results demonstrate that both SSRM and SCM techniques are enabling tools for quantitatively profiling 2D doping/charge carrier distribution in semiconductor devices and SVM is capable of measuring the voltage profile across multiple semiconductor thin layers (as thin as a few atomic layers) with excellent quantitative accuracy (down to microvolts).

The doping profile across the active region of a terahertz QCL device (V843) has been measured by using SSRM and SCM. Individual quantum cascade modules (~37.5 nm in thickness) in the device can be delineated in high-resolution SSRM and SCM images, thanks to the sufficient signal contrast resulted from the δ-doping profile in each module. The dopant concentration of the n-doped active region in V843 is obtained from SSRM and SCM measurements, with the assistance of MBE-grown GaAs staircase-doped structures for calibration. The measured results are in reasonably good agreement with the average designed doping concentration of the device.

The stored data in a Texas Instruments 512 KB embedded NOR flash memory devices with 0.22 μm gate length FGTs were successfully read back using the SCM technique. The SCM signal exhibit sufficient contrast to differentiate the FGs with stored dynamic charge carriers (corresponding to ON state or “1b”) from those without charge carriers (corresponding to OFF state or “0b”). SCM signal also reveals that the quantity of the dynamic charge carrier density could vary significantly from one FG to another one, even though both are at nominally the same ON state. The experimental results demonstrate that the SCM is a reliable approach for reading back the stored data from memory devices.

Different from SSRM and SCM, SVM scans have to be performed on a device that is actively biased. It probes electric potential at nanometer scales and reveals nanoscopic reasons for macroscopic device performance. SVM is applied to two emerging photonic devices: terahertz QCLs and mid-infrared ICLs.

The first application of SVM is to explore the EFD hypothesis in the active region of a terahertz QCL under forward biases. The SVM results provide the direct experimental evidence, showing the formation and evolution of EFDs at different device biases. The higher-field EFD expands linearly with the increase of the externally applied device bias at the expense of the lower-field EFD. The SVM results also reveal that the EFD boundary hops over one complete quantum cascade module every time the device bias increases by ~30 mV, flipping one quantum cascade module entirely from the lower EFD to the higher EFD. The experimental study indicates the importance of designing and growing the quantum-well active region for more uniform and stable electric field profiles, which could lead to new breakthroughs in developing terahertz QCLs with higher performance, such as the long-desired room-temperature operation.

When probing an operating ICL by using SVM, the internal electric potential profile, electric field profile, and net charge carrier density profiles have been directly measured or subsequently derived from the experimental data at nanometer scales. The SVM results convincingly confirm the overall rebalancing of electrons and holes in the active region of the ICL with heavily n-doped profile, even though spatial segregation of electrons and holes has been observed in quantum wells of the active section due to the unique type-II band-gap alignment across the InAs-GaInAs interface. The clamping or non-clamping of charge carrier density in the active region are verified from SVM data in the presence or absence of stimulated emission, respectively. These advanced SPM techniques connect internal mechanisms with external measures. They will find a wide range of applications in developing new nanoelectronic devices, quantum devices, and optoelectronic devices as well as analyzing device failures and performance degradation. In combination with new nanoprobes [72], the SPM is expected to resolve inner workings at subnanometer scales or even at atomic scales.


Corresponding author: Dayan Ban, Waterloo Institute for Nanotechnology, Department of Electrical and Computer Engineering, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, Canada N2L 3G1, e-mail:

Acknowledgments

We acknowledge financial support from Natural Science and Engineering Research Council (NSERC) of Canada, Canadian Foundation of Innovation (CFI), the CMC Microsystems, and Ontario Research Fund (ORF). The authors would like to thank valuable discussions and technical support from Prof. Rui Q. Yang, Dr. Emmanuel Dupont, and Mr. Sylvain Laframboise.

References

[1] Binnig G, Quate CF, Geber Ch. Atomic force microscope. Phys. Rev. Lett. 1986, 56, 930–933.10.1103/PhysRevLett.56.930Suche in Google Scholar

[2] De Wolf P, Snauwaert J, Clarysse T, Vandervorst W, Hellemans L. Characterization of a point-contact on silicon using force microscopy-supported resistance measurements. Appl. Phys. Lett. 1995, 66, 1530–1532.10.1063/1.113636Suche in Google Scholar

[3] De Wolf P, Stephenson R, Trenkler T, Clarysse T, Hantschel T, Vandervorst W. Status and review of two-dimensional carrier and dopant profiling using scanning probe microscopy. J. Vac. Sci. Technol. B 2000, 18, 361–368.10.1116/1.591198Suche in Google Scholar

[4] Dixon-Warren StJ, Lu RP, Ingrey S, Macquistan D, Bryskiewicz T, Smith G, Bryskiewicz B. Scanning spreading resistance microscopy study of a metalorganic chemical vapor deposited grown InP optoelectronic structure. J. Vac. Sci. Technol. A2001, 19, 1752–1757.10.1116/1.1366704Suche in Google Scholar

[5] Edwards H, MiGlothlin R, Martin RS, Elisa U, Gribelyuk M, Mahaffy R, Shih CK, List RS, Ukraintsev VA. Scanning capacitance spectroscopy: an analytical technique for pn-junction delineation in Si devices. Appl. Phys. Lett. 1998, 72, 698–700.10.1063/1.120849Suche in Google Scholar

[6] Kuntze SB, Ban D, Sargent EH, Dixon-Warren StJ, White KJ, Hinzer K. Electrical scanning probe microscopy: investigating the inner workings of electronic devices. J. Crit. Rev. Solid State Mater. Sci. 2005, 30, 71–124.10.1080/10408430590952523Suche in Google Scholar

[7] Tsuji T, Yonezu H, Ohshima N. Reduction of surface roughness of an AlAs/GaAs distributed Bragg reflector grown on Si with strained short-period superlattices. J. Cryst. Growth 1999, 201–202, 1010–1014.10.1016/S0022-0248(99)00031-7Suche in Google Scholar

[8] Behrend J, Carlin JF, Sirbu AV, Berseth CA, Rudra A, Kapon E. InAsP/InGaAsP periodic gain structure for 1.5 μm vertical cavity surface emitting laser applications. J. Cryst. Growth 1998, 188, 295–299.10.1016/S0022-0248(98)00084-0Suche in Google Scholar

[9] De Wolf P, Vandervorst W, Smith H, Khalil N. Comparison of two-dimensional carrier profiles in metal-oxide-semiconductor field-effect transistor structures obtained with scanning spreading resistance microscopy and inverse modeling. J. Vac. Sci. Technol. B 2000, 18, 540–544.10.1116/1.591228Suche in Google Scholar

[10] Lu RP, Kavanagh KL, Dixon-Warren StJ, Spring Thorpe AJ, Hillier G, Macquistan D, Griswold E. Calibrated scanning spreading resistance microscopy profiling of carrier in III–V structures. J. Vac. Sci. Technol. B 2001, 19, 1662–1670.10.1116/1.1387458Suche in Google Scholar

[11] Kopanski JJ, Marchiando JF, Lowney JR. Scanning capacitance microscopy measurements and modeling: progress towards dopant profiling of silicon. J. Vac. Sci. Technol. B 1996, 14, 242–247.10.1116/1.588455Suche in Google Scholar

[12] Ban D, Sargent EH, Dixon-Warren StJ, Grevatt T, Knight G, Pakulski G, Spring Thorpe AJ, Streater R, White JK. Two-dimensional profiling of carriers in a buried heterostructure multi-quantum-well laser: calibrated scanning spreading resistance microscopy and scanning capacitance microscopy. J. Vac. Sci. Technol. B 2002, 20, 2126–2132.10.1116/1.1511211Suche in Google Scholar

[13] Hammar M, Messmer ER, Luzuy M, Anand S, Lourdudoss S, Landgren G. Topography dependent doping distribution in selectively regrown InP studied by scanning capacitance microscopy. Appl. Phys. Lett. 1998, 72, 815–817.10.1063/1.120902Suche in Google Scholar

[14] De Wolf P, Geva M, Reynolds CL, Hantschel T, Vandervorst W, Bylsma RB. Two-dimensional carrier profiling of InP-based structures using scanning spreading resistance microscopy. J. Vac. Sci. Technol. A 1999, 17, 1285–1288.10.1116/1.581809Suche in Google Scholar

[15] Trenkler T, De Wolf P, Vandervorst W, Hellemans L. Nanopotentiometry: local potential measurements in complementary metal-oxide-semiconductor transistors using atomic force microscopy. J. Vac. Sci. Technol. B 1998, 16, 367–372.10.1116/1.589812Suche in Google Scholar

[16] Ban D, Sargent EH, Dixon-Warren StJ, Calder I, Spring Thorpe AJ, Dworschak R, Este G, White JK. Direct imaging of the depletion region of an InP pn junction under bias using scanning voltage microscopy. Appl. Phys. Lett. 2002, 81, 5057–5059.10.1063/1.1528277Suche in Google Scholar

[17] Ban D, Sargent EH, Dixon-Warren StJ, Letal G, Hinzer K, SpringThorpe AJ, Knight G, White JK. Scanning voltage microscopy on buried heterostructure multi-quantum-well lasers: identification of a diode current leakage path. IEEE J. Quant. Electron. 2004, 40, 118–122.10.1109/JQE.2003.821539Suche in Google Scholar

[18] Ban D, Sargent EH, Hinzer K, Dixon-Warren StJ, Calder I, SpringThorpe AJ, White JK. Direct observation of lateral current spreading in ridge-waveguide lasers using scanning voltage microscopy. Appl. Phys. Lett. 2003, 82, 4166–4168.10.1063/1.1581982Suche in Google Scholar

[19] Ban D, Sargent EH, Dixon-Warren StJ, Hinzer K, Calder I, SpringThorpe AJ, White JK. Scanning voltage microscopy on active semiconductor lasers: the impact of doping profile near an epitaxial growth interface on series resistance. IEEE J. Quant. Electron. 2004, 40, 651–655.10.1109/JQE.2004.828262Suche in Google Scholar

[20] Nardi CD, Desplats R, Perdu P, Guerin C, Beaudoin F, Gauffier JL. Oxide charge measurements in EEPROM devices. Microelectron. Reliab. 2005, 45, 1514–1519.10.1016/j.microrel.2005.07.055Suche in Google Scholar

[21] Bidani L, Baharav O, Sinvani M, Zalevsky Z. Usage of laser timing probe for sensing of programmed charges in EEPROM devices. IEEE Trans. Device Mater. Reliab. 2014, 14.1, 304–310.10.1109/TDMR.2013.2272218Suche in Google Scholar

[22] Köhler R, Tredicucci A, Beltram F, Beere HE, Linfield EH, Davies AG, Ritchie DA, Iotti RC, Rossi F. Terahertz semiconductor-heterostructure laser. Nature 2002, 417, 56–159.10.1109/CLEO.2002.1034500Suche in Google Scholar

[23] Scalari G, Walther C, Fischer M, Terazzi R, Beere H, Ritchie D, Faist J. THz and sub-THz quantum cascade lasers. Laser Photon. Rev. 2009, 3, 45–66.10.1002/lpor.200810030Suche in Google Scholar

[24] Williams BS. Terahertz quantum-cascade lasers. Nat. Photon 2007, 1, 517–525.10.1038/nphoton.2007.166Suche in Google Scholar

[25] Razavipour SG, Dupont E, Chan CWI, Lindskog M, Wasilewski ZR, Xu C, Laframboise SR, Wacker A, Hu Q, Ban D. A wideband high carrier injection terahertz quantum cascade laser based on indirect pumped scheme. Appl. Phys. Lett. 2014, 104, 041111–1-5.10.1063/1.4862177Suche in Google Scholar

[26] Xu C, Razavipour SG, Ban D. Impacts of side strips and ridge width on terahertz quantum cascade lasers with metal-metal waveguides. Opt. Express 2013, 31, 31951–31959.10.1364/CLEO_QELS.2013.JTh2A.60Suche in Google Scholar

[27] Fathololoumi S, Dupont E, Wasilewski ZR, Chan CWI, Razavipour SG, Laframboise SR, Huang S, Hu Q, Ban D, Liu HC. Effect of oscillator strength and intermediate resonance on the performance of resonant phonon-based terahertz quantum cascade lasers. J. Appl. Phys. 2013, 113, 113109–1-17.10.1063/1.4795614Suche in Google Scholar

[28] Razavipour SG, Dupont E, Fathololoumi S, Chan CWI, Lindskog M, Wasilewski ZR, Aers G, Laframboise SR, Wacker A, Hu Q, Ban D, Liu HC. An indirectly pumped terahertz quantum cascade laser with low injection coupling strength operating above 150 K. J. Appl. Phys. 2013, 113, 203107–1-14.10.1063/1.4807580Suche in Google Scholar

[29] Dupont E, Fathololoumi S, Wasilewski ZR, Aers G, Laframboise SR, Lindskog M, Wacker A, Ban D, Liu HC. A phonon scattering assisted injection and extraction based terahertz quantum cascade laser. J. Appl. Phys. 2012, 111, 073111–1-10.10.1063/1.3702571Suche in Google Scholar

[30] Fathololoumi S, Dupont E, Chan CWI, Wasilewski ZR, Laframboise SR, Ban D, Matyas A, Jirauschek C, Hu Q, Liu HC. Terahertz quantum cascade lasers operating up to ~200K with optimized oscillator strength and improved injection tunneling. Opt. Express 2012, 20, 3866–3876.10.1364/OE.20.003866Suche in Google Scholar PubMed

[31] Fathololoumi S, Dupont E, Razavipour G, Laframboise SR, Wasilewski Z, Liu HC, Ban D. On metal contacts of terahertz quantum-cascade lasers with a metal-metal waveguide. Semicond. Sci. Technol. 2011, 26, 105021–1-5.10.1088/0268-1242/26/10/105021Suche in Google Scholar

[32] Fathololoumi S, Dupont E, Razavipour G, Laframboise SR, Wasilewski ZR, Bezinger A, Rafi GZ, Safavi-Naeini S, Ban D, Liu HC. Electrically controlling transverse modes in high power THz quantum cascade lasers. Opt. Express 2010, 18, 10036–10048.10.1364/OE.18.010036Suche in Google Scholar PubMed

[33] Fathololoumi S, Dupont E, Ban D, Graf M, Laframboise SR, Wasilewski ZR, Liu HC. Time resolved thermal quenching of THz quantum cascade lasers. IEEE J. Quant. Electron. 2010, 46, 396–404.10.1109/JQE.2009.2031250Suche in Google Scholar

[34] Fathololoumi S, Ban D, Luo H, Dupont E, Laframboise SR, Boucherif A, Liu HC. Thermal behavior investigation of terahertz quantum cascade lasers. IEEE J. Quant. Electron. 2008, 44, 1139–1144.10.1109/JQE.2008.2001901Suche in Google Scholar

[35] Liu HC, Wächter M, Ban D, Wasilewski ZR, Buchanan M, Aers GC, Cao JC, Feng SL, Williams BS, Hu Q. Effect of doping concentration on the performance of terahertz quantum-cascade lasers. Appl. Phys. Lett. 2005, 87, 141102–1-3.10.1063/1.2067699Suche in Google Scholar

[36] Yang RQ. Infrared laser based on intersubband transitions in quantum wells. Superlatt. Microstruct. 1995, 17, 77–83.10.1006/spmi.1995.1017Suche in Google Scholar

[37] Kim M, Canedy CL, Bewley WW, Kim CS, Lindle JR, Abell J, Vurgaftman I, Meyer JR. Interband cascade laser emitting at λ=3.75 μm in continuous wave above room temperature. Appl. Phys. Lett. 2008, 92, 191119–1-3.Suche in Google Scholar

[38] Yang RQ. Novel concepts and structures for infrared lasers. Chap. 2. In Long Wavelength Infrared Emitters Based on Quantum Wells and Superlattices, Helm, M. Ed., Gordon & Breach Publisher: Singapore, 2000.Suche in Google Scholar

[39] Vurgaftman I, Bewley WW, Canedy CL, Kim CS, Kim M, Merritt CD, Abell J, Lindle JR, Meyer JR. Rebalancing of internally generated carriers for mid-infrared interband cascade lasers with very low power consumption. Nat. Commun. 2011, 2, 585–1-7.10.1038/ncomms1595Suche in Google Scholar PubMed

[40] Bewley WW, Canedy CL, Kim CS, Kim M, Merritt CD, Abell J, Vurgaftman I, Meyer JR. High-power room-temperature continuous-wave mid-infrared interband cascade lasers. Opt. Express 2012, 20, 20894–20901.10.1364/OE.20.020894Suche in Google Scholar PubMed

[41] Bewley WW, Canedy CL, Kim CS, Kim M, Merritt CD, Abell J, Vurgaftman I, Meyer JR. Continuous-wave interband cascade lasers operating above room temperature at λ=4.7–5.6 μm. Opt. Express 2012, 20, 3235–3240.10.1364/OE.20.003235Suche in Google Scholar PubMed

[42] Dhar RS, Campbell J, Kawaliye MA, Green M, Ban D. Direct charge measurements to read back stored data in nonvolatile memory devices using scanning capacitance microscopy. J. Vac. Sci. Technol. B 2013, 31, 061801–1-6.10.1116/1.4825200Suche in Google Scholar

[43] Dhar RS, Ban D. Determination of doping concentration in THz QCL device using calibrated scanning spreading resistance microscopy and scanning capacitance microscopy. J. Microsc. 2013, 251, 35–44.10.1111/jmi.12040Suche in Google Scholar PubMed

[44] Dhar RS, Razavipour SG, Dupont E, Xu C, Laframboise S, Wasilewski Z, Hu Q, Ban D. Direct nanoscale imaging of evolving electric field domains in quantum structures. Sci. Rep. 2014, 4, 7183–1-9.10.1038/srep07183Suche in Google Scholar PubMed PubMed Central

[45] Dhar RS, Li L, Ye H, Razavipour SG, Wang X, Yang RQ, Ban D. Nanoscopically resolved dynamic charge-carrier distribution in operating interband cascade lasers. Laser Photon. Rev. 2015, 9, 224–230.10.1002/lpor.201400143Suche in Google Scholar

[46] Ban D. Direct nanometric probing of the inner workings of lasing optoelectronic devices. PhD dissertation, University of Toronto, 2003.Suche in Google Scholar

[47] Jiang Y, Li L, Tian Z, Ye H, Zhao L, Yang RQ, Mishima TD, Santos MB, Johnson MB, Mansour K. Electrically widely tunable interband cascade lasers. J. Appl. Phys. 2014, 115, 113101–1-9.10.1063/1.4865941Suche in Google Scholar

[48] Nardi CD, Desplats R, Perdu P, Guerin C, Gauffier JL, Amundsen TB. Direct measurements of charge in FG transistor channels of flash memories using scanning capacitance microscopy. In Proceedings of the 32nd International Symposium for Testing and Failure Analysis (ISTFA), 2006, p. 86.Suche in Google Scholar

[49] Deleruyelle D, Molas G, De Salvo B, Gely M, Lafond D. Single-electron phenomena in ultra-scaled floating-gate devices and their impact on electrical characteristics. In Proceedings of the 1st International Conference on Memory Technology and Design (ICMTD), France, 2005, p. 25.10.1016/j.sse.2005.10.023Suche in Google Scholar

[50] De Wolf P, Clarysse T, Vandervorst W, Snauwaert J, Hellemans L. One- and two-dimensional carrier profiling in semiconductors by nanospreading resistance profiling. J. Vac. Sci. Technol. B 1996, 14, 380–385.10.1116/1.588478Suche in Google Scholar

[51] De Wolf P, Clarysse T, Vandervorst W. Quantification of nanospreading resistance profiling data. J. Vac. Sci. Technol. B 1997, 16, 320–326.10.1116/1.589804Suche in Google Scholar

[52] Stephenson R, Verhulst A, De Wolf P, Caymax M, Vandervorst W. Contrast reversal in scanning capacitance microscopy imaging. Appl. Phys. Lett. 1998, 73, 2597–2599.10.1063/1.122517Suche in Google Scholar

[53] Kang ES, Kang JW, Hwang HJ, Lee JH. Nondestructive one-dimensional scanning capacitance microscope dopant profile determination method and its application to three-dimensional dopant profiles. J. Vac. Sci. Technol. A 2000, 18, 1338–1344.10.1116/1.582473Suche in Google Scholar

[54] Zavyalov VV, McMurray JS, Williams CC. Scanning capacitance microscope methodology for quantitative analysis of p-n junctions. J. Appl. Phys. 1999, 85, 7774–7783.10.1063/1.370584Suche in Google Scholar

[55] Kopanski JJ, Marchiando JF, Rennex BG. Carrier concentration dependence of the scanning capacitance microscopy signal in the vicinity of p-n junctions. J. Vac. Sci. Technol. B 2000, 18, 409–413.10.1116/1.591243Suche in Google Scholar

[56] Kopanski JJ, Marchiando JF, Berning DW, Alvis R, Smith HE. Scanning capacitance microscopy measurement of two-dimensional dopant profiles across junctions. J. Vac. Sci. Technol. B 1998, 16, 339–343.10.1116/1.589807Suche in Google Scholar

[57] Dhar RS, Dixon-Warren StJ, Kawaliye MA, Campbell J, Green M, Ban D. Readback stored charges on nonvolatile memory devices. In MRS Fall 2012 Conference, 2012, p. 1527.10.1557/opl.2013.416Suche in Google Scholar

[58] Nardi CD, Desplats R, Perdu P, Guérin C, Gauffier JL, Amundsen TB. EEPROM failure analysis methodology: can programmed charges be measured directly by electrical techniques of scanning probe microscopy? In Proceedings of the 31st International Symposium for Testing and Failure Analysis (ISTFA), 2005, pp. 256–261.Suche in Google Scholar

[59] Takasaki Y, Yamamoto T. Cross-section analysis of electric devices by scanning capacitance microscope. Microelectron. Reliab. 1999, 39, 987–990.10.1016/S0026-2714(99)00135-3Suche in Google Scholar

[60] Born A, Wiesendanger R. Guidelines for two-dimension dopant profiling using SCM. In Proceedings of 26th International Symposium for Testing and Failure Analysis (ISTFA), 2000, pp. 521–527.10.31399/asm.cp.istfa2000p0521Suche in Google Scholar

[61] V. Metrology Group, Application Modules: Dimension and Multimode Manual. Software Version 5.12r4. Revision B, 2003.Suche in Google Scholar

[62] Erickson AN. The role of the atomic force microscope in failure and yield analysis. In Tutorial of the 27th International Symposium for Testing and Failure Analysis (ISTFA), 2001.Suche in Google Scholar

[63] Choi KK, Levine BF, Malik RJ, Walker J, Bethea CG. Periodic negative conductance by sequential resonant tunneling through an expanding high-field superlattice domain. Phys. Rev. B 1987, 35, 4172–4175.10.1103/PhysRevB.35.4172Suche in Google Scholar PubMed

[64] Grahn HT, Schneider H, Klitzing KV. Optical detection of high field domains in GaAs/AlAs superlattices. Appl. Phys. Lett. 1989, 54, 1757–1759.10.1063/1.101282Suche in Google Scholar

[65] Grahn HT, Haug RJ, Muller W, Ploog K. Electric-field domains in semiconductor superlattices: a novel system for tunneling between 2D systems. Phys. Rev. Lett. 1991, 67, 1618–1621.10.1103/PhysRevLett.67.1618Suche in Google Scholar PubMed

[66] Grahn HT, Schneider H, Von Klitzing K. Optical studies of electric field domains in GaAs-Alx Ga1-x As superlattices. Phys. Rev. B 1990, 41, 2890–2899.10.1103/PhysRevB.41.2890Suche in Google Scholar PubMed

[67] Wienold M, Schrottke L, Giehler M, Hey R, Grahn HT. Nonlinear transport in quantum-cascade lasers: the role of electric-field domain formation for the laser characteristics. J. Appl. Phys. 2011, 109, 073112–1-6.10.1063/1.3573504Suche in Google Scholar

[68] Vurgaftman I, Bewley WW, Canedy CL, Kim CS, Kim M, Ryan Lindle J, Merritt CD, Abell J, Meyer JR. Mid-IR type-II interband cascade lasers. IEEE J. Select. Top. Quant. Electron. 2011, 17, 1435–1444.10.1109/JSTQE.2011.2114331Suche in Google Scholar

[69] Agrawal GP, Dutta NK. Long Wavelength Semiconductor Lasers, Van Nostrand Reinhold Co. Inc., New York, 1986.10.1007/978-94-011-6994-3Suche in Google Scholar

[70] Bewley WW, Lindle JR, Canedy CL, Kim M, Kim CS, Larrabee DC, Vurgaftman I, Meyer JR. Gain, loss, and internal efficiency in interband cascade lasers emitting at λ=3.6–4.1 μm. J. Appl. Phys. 2008, 103, 013114–1-4.10.1063/1.2831225Suche in Google Scholar

[71] Soibel A, Kamjou M, Yueming Q, Hill CJ, Yang RQ. Optical gain, loss, and transparency current in high performance mid-infrared interband cascade lasers. J. Appl. Phys. 2007, 101, 093104–1-4.10.1063/1.2723188Suche in Google Scholar

[72] Rezeq M, Ali A, Barada H. Fabrication of nano ion-electron sources and nano-probes by local electron bombardment. Appl. Surf. Sci. 2015, 333, 104–109.10.1016/j.apsusc.2015.02.006Suche in Google Scholar

Received: 2015-5-31
Accepted: 2015-7-24
Published Online: 2015-10-27
Published in Print: 2016-6-1

©2016 by De Gruyter

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Heruntergeladen am 17.12.2025 von https://www.degruyterbrill.com/document/doi/10.1515/ntrev-2015-0035/html
Button zum nach oben scrollen