You are on page 1of 7

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 61, NO.

6, JUNE 2014 1929

Integration of a Piezoelectric Layer on Si FinFETs


for Tunable Strained Device Applications
Buket Kaleli, Raymond J. E. Hueting, Senior Member, IEEE, Minh D. Nguyen, and Rob A. M. Wolters

Abstract— Earlier theoretical reports predicted that the usage


of a piezoelectric stressor layer around the FinFET, i.e., the
PiezoFET, offers a great potential for steep subthreshold slope
devices. For the first time, we analyzed the practical realization
of such PiezoFETs comprising a piezoelectric stressor layer, lead–
zirconate–titanate (PZT), and aluminum–nitride (AlN) deposited
on n-type silicon FinFETs. A high-piezoelectric response in the
range of 100 pm/V has been obtained for the PZT PiezoFET
evidencing the converse piezoelectric effect in the device. The
piezoelectric response for the AlN device was much less (13 pm/V)
as expected. Underlying device properties, such as subthreshold
swing (SS) and low-field electron mobility have been significantly Fig. 1. PiezoFET schematic cross section with narrow fins illustrating the
affected by the presence of the PZT stressor. A 20%–50% change formed strain condition corresponding to an upward piezoelectric response.
in the mobility and a change in the SS (about 5 mV/decade) have The white arrows show the PZT deformation, while the black arrows show
been observed. The change can be attributed to the strain induced the corresponding formed stress in the silicon fins. A schematic view of the
reduction of the interface trap density at the Si/SiO2 interface. layer stack along a vertical line through an Si fin is shown on the left. The
figure is not to scale.
This strain is partly formed by the bias over the piezoelectric
layer, which indicates the converse piezoelectric effect related
tunable strain in both the silicon channel and gate oxide. that, it is possible to obtain a device with a subthreshold
Index Terms— Aluminum–nitride (AlN), FinFET, lead– swing (SS) below the fundamental limit of 60 mV/decade
zirconate–titanate (PZT), piezoelectric effect, strain, stress. at room temperature [8]. Newns et al. [9] proposed the
piezoelectric transistor (PET) based on the piezoresistance
I. I NTRODUCTION effect on special channel materials manifesting a pressure-
induced metal–insulator transition. Their calculations show
F ERROELECTRIC and piezoelectric materials find appli-
cation in high-density capacitors [1], sensors, microme-
chanical systems [2], [3], and nonvolatile memories [4], [5].
that the PETs performance outpaces the conventional FETs
with respect to the lower supply voltage and higher switching
Recently, there has been a growing number of reports on speed. In [10] and [11], it has been theoretically predicted
novel steep-subthreshold devices where ferroelectric and/or that the piezoelectric materials can be utilized in standard
piezoelectric thin films are employed. For instance, Salahuddin semiconductor materials, such as silicon (Si) or germanium,
and Datta [6] proposed to use ferroelectrics as a gate dielectric for obtaining steep-subthreshold slope switches.
to obtain a negative capacitance for improvement of the In this paper, the (converse) piezoelectric effect of a ferro-
device properties. Later on, Rusu et al. [7] demonstrated this electric lead–zirconate–titanate (Pb[Zrx Ti1−x ]O3 or PZT) thin
concept experimentally on p-MOSFET devices. They show film is experimentally studied as a channel stressor in the Si
FinFET. We call this novel device structure the PiezoFET.
Manuscript received November 2, 2013; revised January 30, 2014; accepted The motivation of using a piezoelectric material as a stressor
March 31, 2014. Date of publication May 1, 2014; date of current version for FinFETs is to increase the mobility and to reduce the
May 16, 2014. This work was supported in part by the Dutch Technology
Foundation STW, in part by an Applied Science Division, NWO, and in part SS caused by band-offset modification [10], [11].
by the Technology Program, Ministry of Economic Affairs. The review of The PiezoFET is basically a four terminal device consisting
this paper was arranged by Editor J. C. S. Woo. of two separate gates (Fig. 1): 1) the conventional gate that
B. Kaleli and R. J. E. Hueting are with the MESA+ Institute for
Nanotechnology, University of Twente, Enschede 7522, The Netherlands electrostatically controls the channel, which we call the bottom
(e-mail: b.kaleli@utwente.nl; r.j.e.hueting@utwente.nl). gate (BG) and 2) a top gate (TG or pigate) to control the
M. D. Nguyen is with the MESA+ Institute for Nanotechnology, University channel strain. The PZT material was chosen as a stressor
of Twente, Enschede 7522, The Netherlands, Solmates B.V., Enschede 7522
NB, The Netherlands, and also with the International Training Institute for because of its higher piezoelectric response among the other
Materials Science, Hanoi University of Science and Technology, Hanoi 10000, piezoelectric materials [3]. In addition, we have investigated
Vietnam (e-mail: d.m.nguyen@utwente.nl). the effect of aluminum–nitride (AlN) film for comparison.
R. A. M. Wolters is with the MESA+ Institute for Nanotechnology,
University of Twente, Enschede 7522, The Netherlands, and also with This paper is outlined as follows. In Section II, we describe
NXP Semiconductors, Eindhoven 5656 AE, The Netherlands (e-mail: the device fabrication followed by the piezoelectric response
r.a.m.wolters@utwente.nl). in Section III. In Section IV, we report on the electrical
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org. characterization and the analysis of the device properties.
Digital Object Identifier 10.1109/TED.2014.2316164 Finally, conclusions are drawn in Section V.
0018-9383 © 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
1930 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 61, NO. 6, JUNE 2014

III. P IEZOELECTRIC R ESPONSE


Fig. 1 shows a schematic cross section of the fabricated
PiezoFET presenting two Si fins under applied bias across a
homogenously deposited PZT (or AlN) layer. The displace-
ments are indicated here when the PZT is elongated in the
direction of the electric field between BG and TG. The white
arrows indicate the piezoelectric displacement in the PZT
layer, while the black arrows indicate the corresponding strain
formed inside the Si fins. This happens for a negative bias over
Fig. 2. Optical micrograph of a typical 5-fin PiezoFET. The bottom gate,
top gate, drain, and source are indicated by BG, TG, D, and S, respectively.
the PZT layer (Vπ = VTG − VBG < 0 V). In this case, the
stress formed in the Si fin can be described as follows: 1) due
to the PZT layer at the fin sidewall region, the Si fin is under
biaxial tensile stress along x (fin width) and z (fin height), and
under uniaxial compressive stress along y (channel length);
2) due to the PZT layer on top of the fin, biaxial compressive
stress along x and y (channel length) is formed; and 3) due to
the layer on the BOX region, tensile stress in Si along x and
compressive stress along y is formed.
The overall strain effect depends on the fin width and aspect
ratio. For narrow fins condition 1) dominates, whereas for
wider structures condition 2) is more important. In practice,
Fig. 3. HR-TEM images of a magnified image of one fin of a PiezoFET condition 3) would not be important for narrow fins, as is
with (a) 100 nm and (b) 20-nm-wide fin and 10-nm LNO + 100-nm PZT
layer stack.
explained in Section IV. It should be noted that for each fin
width, the net stress depends on the amount of the individual
II. D EVICE FABRICATION stress components. Fig. 1 is shown for a narrow fin structure
Devices were fabricated on a (001)-oriented undoped where the PZT on the BOX layer (and at the fin sidewalls)
silicon-on-insulator (SOI) wafer with a buried-oxide (BOX) determines the net component. For downward deflection [at
layer of 1-μm thickness. The 1.5-μm-thick Si device layer was positive bias over the PZT layer (Vπ = VTG − VBG > 0 V)],
thinned down to 150 nm by thermal oxidation. The (110)/[110] all the stress components change direction. Note that in this
oriented Si fins with different fin widths (WFIN = 15–1000 nm) discussion, a possible stress relieving effect of the intermediate
were patterned with e-beam lithography. After dry etching layers was not considered.
of the Si fins, a 11-nm-thick gate oxide layer was thermally A laser Doppler vibrometer (LDV) measurement was per-
grown. A 6-nm-thick TiN and 12-nm-thick α-Si deposition formed to characterize the piezoelectric response of the
was done in our in-house built cluster tool using atomic-layer PZT layer on our FinFET devices. A sinusoidal ac-voltage
deposition (ALD) and chemical vapor deposition techniques, of 1 V and at 8-kHz frequency was applied here to a
respectively [12]. Following the BG stack patterning, As+ ion 100-nm-thick PZT layer. The response under these measure-
implantation (1015-cm−2 dose, 50 keV), and a rapid thermal ment conditions is shown in Fig. 4(a) and (b) for FinFETs
annealing (900 °C, 30 s) were performed. Afterwards, a stack with 30- and 1000-nm fin width, respectively. It is clear that
containing a PZT film and an LaNiO3 (LNO) buffer layer was the vertical displacement is enhanced on top of the fins and
deposited using a SolMateS large area pulsed-laser deposition there is a converse piezoelectric effect. The displacement is
(PLD) system at 650 °C. PiezoFETs containing sputtered AlN also high outside the fin area, under the TG. Fig. 4(c) and (d)
were also processed. Patterning of the piezoelectric layers was shows the line scan of the displacement across the dotted
performed by wet etching. Finally, the Pt/Ti TG and contact lines in Fig. 4(a) and (b). A piezoelectric coefficient of about
pads were defined by a lift off process. The PiezoFET is shown d33 , f = 100–110 pm/V on the fin area and about d33, f =
in the optical micrograph in Fig. 2. There is a 2 μm overlap 80–90 pm/V outside the fin area above the top contact has
between the TG and the piezoelectric layer. been determined. As will be discussed in Section IV, the
Fig. 3(a) and (b) shows TEM pictures of a 100- and actual voltage drop over the PZT layer is less than the applied
20-nm-wide fin PiezoFET with a 10-nm-thick LNO and voltage. The values are doubled when the applied nominal
100-nm-thick PZT layer. The TG is not visible in the figures voltage was increased to 2 V. This is an indication that these
due to the sample preparation (i.e., Pt protective layer). In these values are real displacements.
images, good step coverage of the TiN as well as poly-Si is The enhancement of displacement on the channel region
recognized, whereas the step coverage of the LNO/PZT stack could be due to the path length difference of the laser beam
is poor. The PZT layer thickness decreases to 20–50 nm, while between the top and the outer side of the fin area. The
the LNO layer thickness decreases to 2–5 nm at the sidewalls. path is shorter (by a fin height amount ≈150 nm) for the
Although the step coverage is poor, electrical measurements beam deflected from the top. Furthermore, there is a 10%
show that there is no direct electrical short circuit between the difference in the maximum displacement between the wide
TiN BG and the Ti/Pt TG in the structures, as will be discussed fin and narrow fin device. The maximum displacement is
in Section IV. enhanced on wide fins possibly due to their relatively wide
KALELI et al.: INTEGRATION OF A PIEZOELECTRIC LAYER ON Si FinFETs 1931

Fig. 4. 2-D upward response scans of a 5-fin PiezoFET with a (a) 30 nm and
(b) 1000-nm fin width and L pigate = 10 μm and L gate = 30 μm. Nominally, Fig. 6. Id –Vgs device characteristics of the PiezoFETs for different Vπ with
a sinusoidal ac voltage of 1 V and at 8-kHz frequency was used. The graphs (a) WFIN = 30 nm and (b) WFIN = 1000 nm.
in (c) and (d) are the line scan profiles of the displacement across the dotted
line in (a) and (b), respectively. The channel response is indicated in between
two vertical dotted lines. A stack of 10-nm LNO + 100-nm PZT thin films
was used in this structure.
devices. Deposition and processing of the LNO/PZT layers
show no electrical and physical degradation of the Si body.
The XPS analysis on our planar devices also did not indicate
any diffusion of atoms between the different material layers.
The observed shift in the I –V curve could be due to stress in
the Si FinFET after the PLD process. It has been reported that
due to the difference in coefficient of thermal expansion (CTE)
of the Si and PZT a thermal stress can be developed in the PZT
film [13]. Since Si has a CTE of 2.6 × 10−6 K−1 [14] and
PZT has a CTE of 6.0 × 10−6 K−1 [15], additional thermal
biaxial in-plane tensile stress is formed in the PZT film during
cooling down process from deposition temperature (650 °C)
to room temperature. This will result in biaxial compressive
stress in the Si fin. Thermal expansion related stress profiles
Fig. 5. Id –Vgs device characteristics of a 5-fin FinFET before and after PZT and analysis have been reported previously [16] and [17].
deposition with L pigate = 10 μm, L gate = 30 μm, and WFIN = 100 nm at However, the shift cannot be explained with the strain effect
Vds = 1 V and room temperature.
in the Si alone since the observed shift of around 0.15 V
is relatively large. We will elaborate on this further in this
section.
fin top surface, which is in the order of beam spot size that
In our analysis for each measured I –V curve, the PZT layer
helps to obtain more signal from the fin top surface.
was biased at a fixed voltage between TG and BG (VBG = Vg )
The same LDV measurements were performed on
to observe the converse piezoelectric effect on the operation
PiezoFETs containing 180-nm-thick AlN layers (not shown)
of the device. Fig. 6 shows the Id –Vgs curves of PiezoFETs
yielding a piezoelectric coefficient of around 13 pm/V. This
with a 30- and 1000-nm fin width.
lower value is expected due to material properties. However,
A threshold shift depending on Vπ as well as higher
it is higher than the piezoelectric coefficient of bulk AlN. This
off current (Ioff ) is observed. Furthermore, unlike reported
is attributed to the mentioned reasons before.
for FerroFETs [7], no hysteresis behavior in our PiezoFETs
is observed. Generally, ferroelectric materials show hystere-
IV. E LECTRICAL C HARACTERIZATION AND A NALYSIS sis due to the polarization effect [18]. However, the rela-
Fig. 5 shows the Id − Vgs device characteristics of the same tively high leakage current interferes with the polarization
100-nm-wide fin device before and after the 100-nm-thick PZT switching in our devices, which results in a nonhysteretic
deposition. A negative shift of threshold voltage (Vth ) is behavior. When carefully observing Fig. 6, it appears that
observed after PZT deposition. This shift is observed in all there is an exponential dependency of Vth on Vπ , reaching
1932 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 61, NO. 6, JUNE 2014

Fig. 7. (a) J –Vπ characteristics of the PZT capacitor on a planar sur-


face and on 5-fin 30, 100, and 1000-nm FinFET with L TG = 10 μm. Fig. 8. SS-Vπ graph of 5-fin device with WFIN = 30, 100, and 1000 nm at
(b) Modified schematic equivalent circuit of the measurement setup utilized Vds = 1 V.
for the PiezoFET I –V measurements. The piezoelectric layer is represented
by a leaky resonator with a capacitance Cpi .

explained by charge trapping at the Si/SiO2 interface-induced


+1.5 and −2 V under −3 and +3 V applied bias over the during the processing [19]. Further, the SS has reduced after
PZT layer, respectively. The results also indicate that there is the PZT deposition for narrow fins. This could indicate that the
an asymmetry regarding the polarity. gate oxide (SiO2 ) layer has been stressed along its interface
Related to this, the leakage current of the PZT of the with the Si body, which results in large structural deformation
PiezoFET shows an exponential dependency with Vπ and an and possibly even affecting the band alignment. The effect of
almost symmetric behavior with the applied bias, see Fig. 7(a). fixed strain on the trapping probability and band modulation
The 5-fin devices with 30-, 100-, and 1000-nm fin width was reported previously [20], [21]. We report this on tunable
show a leakage current within the same range. The planar strain in our devices as well.
capacitor, on the other hand, shows much less leakage. This Since SiO2 is less stiff than Si, most of the strain will be
is an indication of excess leakage due to nonuniform step absorbed in the oxide and possibly at its interface. Earlier, it
coverage of the PZT layer (see Fig. 3). PiezoFETs with a 180- was shown by finite element method (FEM) simulations that
nm-thick AlN layer show very similar characteristics with the higher strain values in the gate dielectric are present compared
same amount of Vth for the 30-nm-wide devices. However, with the Si body [22]. This can result in a reduced effective
for wide fins Vth is smaller due to a lower leakage current interface trapping of charge carriers, hence a lower SS. A part
(data is not shown here). of the threshold voltage shift could also be explained by this
All the measurement data indicate a relation between Vth phenomenon.
and the leakage current formed possibly by a leaky piezo- Up to 4 mV/decade improvement (decrease) in the SS is
electric capacitor Cpi on top of the FET connected by the observed after the PZT deposition on narrow fin devices. For
BG resistance, as shown in a simple electrical dc equivalent the 1000-nm-wide fin device, however, it increases. Decreasing
circuit diagram, see Fig. 7(b), where current distribution the SS for narrow fins indicates a reduction of the interface
effects are neglected. In the ideal case, there is no current trap density. Conversely, for wide fins the SS increases and this
through Cpi and the surface potential in the Si fin equals Vgate . indicates an increase in the interface trap density. Change in
However, when there is a (variable) resistor (Rpi ) parallel to the density of states depending on the fin width is possibly due
the piezoelectric capacitor, the surface potential in the Si fin to different crystal orientation of the fin at the sidewall and the
will be affected. The influence of Rpi on the surface potential surface. In addition, we expect a different stress configuration
is determined by the ratio Rgate /Rpi ; the higher this ratio the for wide and narrow fins due to the dominance of the surface
greater its influence. Our measurement data indicate that Rgate for wide fins and the sidewalls for narrow fins. This would
is mainly determined by the 30 μm × 50 μm external gate have a different effect on the band alignment as well.
connection formed by the 6-nm-thick TiN layer, corresponding Furthermore, the SS is modified (up to 6 mV/decade) by
to a value of 445 . For estimating Rpi , we need to determine biasing the PZT layer up to 3 V. This improvement is due
the leakage current through the PZT layer. For Vπ = 3 V, for to the converse piezoelectric effect. For devices with 30- and
instance, a leakage current of 2.5 mA was measured yielding 100-nm fin width, the SS is decreased with an applied PZT
Rpi = 755 . Using these calculated resistances, a ≈2/3· Vπ bias (Fig. 8). Note that short channel effects become dominant
voltage drop is estimated to be across the PZT layer whereas for a 1000-nm-wide fin device and the conclusions drawn from
the rest of the voltage drop is across the gate. The surface device characteristics will not be accurate.
potential changes by this amount and therefore a threshold The effective electron mobility of the PiezoFET has been
voltage shift occurs. calculated by compensating for the Vth . Fig. 9(a)–(c) shows
To analyze the (converse) piezoelectric effect, the SS-Vπ the effective electron mobility as a function of the electric
graph is shown in Fig. 8. Despite the expected improved elec- field (μeff − E) for our 5-fin (110)/[110] PiezoFETs with
trostatic gate control in the narrow channel undoped FinFETs, WFIN = 30, 100, and 1000 nm under different bias over the
the SS is above 60 mV/decade for all devices. This can be PZT layer. For comparison, the μeff − E curve of the (100)
KALELI et al.: INTEGRATION OF A PIEZOELECTRIC LAYER ON Si FinFETs 1933

Fig. 10. Effective electron mobility-Vπ curves of devices (a) at a


5 × 104 V/cm field for 30- and 100-nm-wide fin devices and at a
1 × 105 V/cm field for a 1000-nm-wide fin device and (b) at a 2 × 105 V/cm
field for 30- and 100-nm-wide fin devices and at a 4 × 105 V/cm field for a
1000-nm-wide fin device.

[Fig. 10(a)] and at strong inversion [Fig. 10(b)]. For the 30-
and 100-nm-wide fin devices, the applied fields are at E = 5 ×
104 and 2 × 105 V/cm for moderate and strong inversion,
respectively. Due to the fact that the WFIN = 1000-nm
device acts as a partially depleted SOI device, a higher
electric field is needed to obtain strong inversion. Therefore,
for a 1000-nm-wide device, the effective mobility-Vπ curve
is plotted at a higher field of 1 × 105 and 4 × 105 V/cm for
moderate and strong inversion, respectively.
At low fields, the mobility of the 30- and 100-nm-wide
devices shows an increasing trend under positive applied bias
and a decreasing trend under negative applied bias. At high
fields, the mobility decreases again under negative applied
bias. But no trend is observed for a positive applied bias. The
next step is to discuss the possible reasons for the observed
mobility behavior.

A. After Deposition of (Before Biasing) the PZT Layer


Fig. 9. Effective electron mobility–electric field curves of PiezoFETs for As shown in Fig. 10, the mobility has increased for 1000-
different biasing over the PZT layer with (a) 5-fins of WFIN = 1000 nm
and (100) MOSFET obtained from Takagi et al. [24], (b) 5-fins of WFIN = and 100-nm-wide fin devices after the LNO/PZT deposition for
100 nm, and (c) 5-fins of WFIN = 30 nm and (110) FinFET with WFIN = Vπ = 0 V. Devices show higher mobility values at lower fields
25 nm obtained from Rudenko et al. [23] and (110) MOSFET obtained from in Fig. 10(a) (in the moderate inversion regime when Coulomb
Takagi et al. [24].
or ion scattering is important) after the PZT deposition. The
trend is the same at high fields, except for the 30-nm-wide
device, which shows a decrease in mobility [Fig. 10(b)].
MOSFET and (110) FinFET, (110) MOSFET are shown in The difference in the effective mobility and the shift in
Fig. 9(a) and (c), respectively [23], [24]. Compared with the the threshold voltage in the I –V curves indicate that strain is
mobility values given in [23], the low-field mobility is higher formed after deposition of the LNO/PZT stack. Furthermore,
whereas the high-field mobility is lower in our devices. The as mentioned earlier, oxide deformation after PZT deposition
higher mobility at lower fields is an indication of less effective possibly changes the interface trap density. At low-fields
Coulomb scattering, most probably because of a higher fin and Coulomb scattering is important and that is affected by the
the use of SiO2 as a gate dielectric in our case. Rudenko et al. traps. Fig. 9 shows that the most significant mobility change
[23] used a high-κ material, in which dipoles that are close to occurs at low fields. Since both the mobility and SS are
the channel could increase the Coulomb scattering. However, affected by the PZT deposition, this strongly indicates that
the lower mobility at higher fields is an indication of more charge trapping is the root cause for the increase in the
surface imperfections in our devices. low-field mobility [25].
Up to ≈20% increase in the mobility was observed for the At high fields, on the other hand, surface roughness scat-
WFIN = 1000-nm device after PZT deposition. The effect on tering is important, which could also be affected by strain.
WFIN = 100- and 30-nm devices at low fields is stronger; As mentioned earlier, we expect a biaxial tensile stress in
yielding up to ≈50% increase in mobility. No significant the PZT due to difference in CTE when no external bias
change is observed at high fields. is applied. The step coverage of the PZT is not as ideal as
To show the trend in the mobility more clearly, the mobility- shown in Fig. 2. Therefore, we expect that the stress profile
Vπ curve has been plotted when biased at moderate inversion is not determined by the PZT layer on the BOX layer, but
1934 IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 61, NO. 6, JUNE 2014

more by the PZT layer located at the fin sidewall. Hence, for field, more tensile stress is formed along the channel length
narrow fins (WFIN << HFIN ), a strain gradient parallel to the direction and compressive stress in fin height direction and the
surface of the (110) fin sidewall is important. In this case, mobility increases [26]. The situation is different for a wide
for narrow fins, biaxial compressive stress along the fin height fin. Biaxial tensile stress at the (001) surface is formed under
and channel length direction is formed. This results in uniaxial positive applied field and this tends to increase the electron
tensile stress along the fin width direction and consequently mobility [28]. On the other hand, a negative bias will yield
increases the high-field electron mobility [26]. almost no change in the mobility.
For narrow fins under both tensile and compressive stress 1) At Low Field: Fig. 10(a) shows an increasing trend of
along the fin height direction, the conduction band valleys the effective mobility of up to 20% at a positive bias for
move down in energy, though the latter would give a larger 30- and 100-nm-wide fins. The trend is the opposite at a
impact both in the band offset and increased mobility [26]. negative applied bias. For the 1000-nm-wide device, however,
This will result in a negative threshold voltage shift after there is almost no change.
deposition [10], [11]. This could cause a relatively small part 2) At High Field: In Fig. 10(b), a smaller decrease in
of the Vth observed in Fig. 5. Compressive stress along the effective mobility at a negative bias is observed. However, for
fin height direction moves the 2 (“light”) electron valleys a positive bias, it is almost constant in contrast to Fig. 10(a).
down in energy. The conductivity effective mass of electrons Therefore, the trend in mobility under applied biases for
reduces, thereby increasing the electron mobility. Since we 30- and 100-nm-wide devices can be mainly explained by
expect compressive stress along the channel length, this stress the data from literature. However, the effect of tunable strain
configuration is present for narrow fins in our devices (for is found to be less significant than the strain formed by the
WFIN = 30 and 100 nm). We can observe it in Fig. 10(a) and CTE difference.
(b), except for 30-nm-wide device at high field. The tunable strain in the channel can be improved in
For wide fins (WFIN  HFIN ) on the other hand, the different ways. Minimizing the interlayer thicknesses, such as
sidewalls can be neglected compared with the top surface those of the TiN and the gate oxide, and replacing them with
of the fin. Therefore, wide fins (WFIN = 1000 nm) are more stiff materials can be an option. In addition, improving
expected to behave like a planar (001) surface [as in a (001) the step coverage of the piezoelectric layer will prevent the
bulk MOSFET]. Since the Si substrate is much thicker than leakage and will enable straining the channel by the PZT on
the Si fin height, the intermediate layers and the PZT layer, the BOX layer. The PZT layer thickness should be reduced
the in-plane biaxial tensile stress formed in the PZT layer as well, keeping the leakage current in mind. This will lower
will be transferred to the underlying layers down to the Si the (CMOS compatible) supply voltage required for device
substrate [17]. Hence, biaxial tensile stress along the fin width operation, which ultimately leads to a higher switching speed
and channel length direction is formed. This configuration (GHz range) [9], [11]. In this paper, we did not employ a
would result in uniaxial compressive stress along the fin height poling process. Poling would also greatly improve the ferro-
direction and causes the 2 (“light”) electron valleys to move and piezoelectric properties of the PZT layer (see [29]) and
down in energy. The conductivity effective mass of electrons consequently the tunable strain.
decreases and thereby electron mobility increases. This can be
observed in Fig. 10(a) and (b) for WFIN = 1000 nm. V. C ONCLUSION
In summary, before any Vπ biasing, the low-field mobility
We have presented, for the first time, the device fabrication,
(moderate inversion) increases possibly by the oxide defor-
characterization, and analysis of a novel four terminal device
mation that reduces the effective interface traps at the Si/SiO2
structure called the PiezoFET. A negative shift is observed in
interface. On the other hand, the CTE difference formed strain
the Id − Vgs curves after deposition of PZT and AlN films.
in Si influences the high-field mobility (strong inversion). For
This effect could be related to strong gate oxide deformation
wide fins, the high-field mobility increases due to biaxial ten-
due to CTE difference formed strain. Further, the effect of the
sile stress along the fin width and channel length direction in
piezoelectric layer on device properties of the transistors was
the (001) surface. For narrow fins, the mobility also increases
studied in terms of SS and mobility. It has been shown that
due to biaxial compressive stress along the fin height and
for moderate inversion (low field) both parameters are affected
channel length direction along the (110) surface of the fin
by the CTE difference formed strain and the converse piezo-
sidewall.
electric effect. This can possibly be attributed by the oxide
deformation that reduces the interface traps at the Si/SiO2
B. Biasing the PZT Layer: Tunable Strain interface. The high-field mobility is increased (up to ≈50%)
caused by the CTE difference formed strain, while a smaller
A varying bias from −3 to 3 V has been applied over the
effect (≈5%) was observed by the converse piezoelectric effect
PZT layer and the mobility has been calculated. According to
depending on the fin width. Our results have demonstrated the
Fig. 1, we expect an upward deflection under negative whereas
feasibility and great potential of the PiezoFET concept.
downward deflection under a positive applied field [27].
Therefore, for a negative field, we expect the narrow Si fin
to be more compressively stressed along the channel length ACKNOWLEDGMENT
direction and tensile stressed along the fin height direction The authors would like to thank A. A. I. Aarnink and
and the mobility decreases. On the other hand, for a positive S. Smits for the guidance of the atomic-layer deposi-
KALELI et al.: INTEGRATION OF A PIEZOELECTRIC LAYER ON Si FinFETs 1935

tion process and measurement setup and R. Keim for the [24] S. Takagi, A. Toriumi, M. Iwase, and H. Tango, “On the universal-
TEM pictures. ity of inversion layer mobility in Si MOSFET’s: Part II—Effects of
surface orientation,” IEEE Trans. Electron Devices, vol. 41, no. 12,
pp. 2363–2368, Dec. 1994.
[25] C. T. Sah, T. H. Ning, and L. L. Tschopp, “The scattering of electrons
R EFERENCES by surface oxide charges and by lattice vibrations at the silicon-silicon
dioxide interface,” Surf. Sci., vol. 32, no. 3, pp. 561–575, 1972.
[1] M. Klee, D. Wissen, W. Keur, R. Kiewitt, D. Bausen, and P. Lok, “Oxide [26] N. Serra and D. Esseni, “Mobility enhancement in strained n-FinFETs:
films for integrated capacitors in thin film functional modules,” in Proc. Basic insight and stress engineering,” IEEE Trans. Electron Devices,
Mater. Res. Soc. Symp., vol. 655. 2001, pp. CC13.1.1–CC13.1.11. vol. 57, no. 2, pp. 482–490, Feb. 2010.
[2] P. Muralt, “Ferroelectric thin films for micro-sensors and actuators: [27] N. Jalili, J. Wagner, and M. Dadfarnia, “A piezoelectric driven ratchet
A review,” J. Micromech. Microeng., vol. 10, no. 2, pp. 136–146, 2000. actuator mechanism with application to automotive engine valves,”
[3] N. Setter et al., “Ferroelectric thin films: Review of materials, Mechatronics, vol. 13, no. 8, pp. 933–956, 2003.
properties, and applications,” J. Appl. Phys., vol. 100, no. 5, [28] T. Skotnicki et al., “Innovative materials, devices, and CMOS technolo-
pp. 051606-1–051606-46, 2006. gies for low-power mobile multimedia,” IEEE Trans. Electron Devices,
[4] C. P. de Araujo, J. D. Cuchiaro, L. D. McMillan, M. Scott, and vol. 55, no. 1, pp. 96–129, Jan. 2008.
J. F. Scott, “Fatigue-free ferroelectric capacitors with platinum elec- [29] M. D. Nguyen, E. Houwman, M. Dekkers, H. N. Vu, and
trodes,” Nature, vol. 374, no. 6523, pp. 627–629, 1995. G. Rijnders, “A fast room-temperature poling process of piezoelec-
[5] J. F. Scott, Ferroelectric Memories. New York, NY, USA: tric Pb(Zr0.45 Ti0.55 )O3 thin films,” Sci. Adv. Mater., vol. 6, no. 2,
Springer-Verlag, 2000. pp. 243–251, 2014.
[6] S. Salahuddin and S. Datta, “Use of negative capacitance to provide
voltage amplification for low power nanoscale devices,” Nanoletters,
vol. 8, no. 2, pp. 405–410, 2008.
[7] A. Rusu, G. A. Salvatore, D. Jimenez, and A. M. Ionescu,
“Metal-ferroelectric-metal-oxide-semiconductor field effect transistor Buket Kaleli received the B.Sc. and M.Sc. degrees
with sub-60 mV/decade subthreshold swing and internal voltage in physics from Middle East Technical University,
amplification,” in Proc. IEDM, 2010, pp. 16.3.1–16.3.4. Ankara, Turkey, in 2007 and 2009, respectively, and
[8] S. M. Sze and K. K. Ng, Physics of Semiconductor Devices, 3rd ed. the Ph.D. degree from the University of Twente,
New York, NY, USA: Wiley, 2007, pp. 158–160. Enschede, The Netherlands, in 2013.
[9] D. Newns, B. Elmegreen, H. Liu, and G. Martyna, “A low-voltage
high-speed electronic switch based on piezoelectric transduction,”
J. Appl. Phys., vol. 111, no. 8, p. 084509, 2012.
[10] T. van Hemert and R. J. E. Hueting, “Active strain modulation in field
effect devices,” in Proc. ESSDERC, 2012, pp. 125–128.
[11] T. van Hemert and R. J. E. Hueting, “Piezoelectric strain modulation in
FETs,” IEEE Trans. Electron Devices, vol. 60, no. 10, pp. 3265–3270,
Oct. 2013.
[12] H. van Bui, A. Y. Kovalgin, and R. A. M. Wolters, “Growth of sub-
nanometer thin continuous TiN films by atomic layer deposition,” ECS Raymond J. E. Hueting (S’94–M’98–SM’06)
J. Solid State Sci. Technol., vol. 1, no. 6, pp. 285–290, 2012. received the M.Sc. (cum laude) and Ph.D. degrees
[13] G. Han et al., “Stress-controlled Pb(Zr0.52 Ti0.48 )O3 thick films by in electrical engineering from the Delft University
thermal expansion mismatch between substrate and Pb(Zr0.52 Ti0.48 )O3 of Technology, Delft, The Netherlands, in 1992 and
film,” J. Appl. Phys., vol. 110, no. 12, p. 124101, 2011. 1997, respectively.
[14] K. G. Lyon, G. L. Salinger, C. A. Swenson, and G. K. White, “Linear He joined the Semiconductor Components Labo-
thermal expansion measurements on silicon from 6 to 340 K,” J. Appl. ratory at the MESA+ Institute for Nanotechnology,
Phys., vol. 48, no. 3, pp. 865–868, 1977. University of Twente, Enschede, The Netherlands,
[15] K. Kurihara, M. Kondo, K. Sato, M. Ishii, N. Wakiya, and K. Shinozak, in 2005.
“Electrooptic properties of epitaxial lead zirconate titanate films on
silicon substrates,” Jpn. J. Appl. Phys., vol. 46, no. 10S, pp. 6929–6932,
2007.
[16] W. H. Xu, D. Lu, and T.-Y. Zhang, “Determination of residual stresses
in Pb(Zr0.53 Ti0.47 )O3 thin films with Raman spectroscopy,” Appl. Phys.
Minh D. Nguyen received the Degree in chemistry,
Lett., vol. 79, no. 25, pp. 4112–4114, 2001.
the master’s degree in materials science, and the
[17] M. H. Zhao, R. Fu, D. Lu, and T.-Y. Zhang, “Critical thickness for
Ph.D. degree in physics from the University of
cracking of Pb(Zr0.53 Ti0.47 )O3 thin films deposited on Pt/Ti/Si(100)
Twente, Enschede, The Netherlands, in 1999, 2001,
substrates,” Acta Mater., vol. 50, no. 17, pp. 4241–4254, 2002.
and 2010, respectively.
[18] S. L. Miller and P. J. McWhorter, “Physics of the ferroelectric non- His current research interests include various
volatile memory field effect transistor,” J. Appl. Phys., vol. 72, no. 12, piezoelectric MEMS devices, where he concen-
p. 5999, 1992. trates on piezoelectric microdiaphragms and micro-
[19] D. K. Schroder, Semiconductor Material and Device Characterization, cantilevers for microfluidics and microbiosensors
3rd ed. New York, NY, USA: Wiley, 2006. applications.
[20] T. B. Hook and T.-P. Ma, “Electron trapping during high-field tunneling
injection in metal-oxide-silicon capacitors: The effect of gate-induced
strain,” J. Appl. Phys., vol. 62, no. 3, pp. 931–938, 1987.
[21] T. Irisawa, T. Numata, N. Hirashita, T. Tezuka, N. Sugiyama, and
S. Takagi, “Physical understanding of strain-induced modulation of gate
Rob A. M. Wolters received the M.Sc. and
oxide reliability in MOSFETs,” IEEE Trans. Electron Devices, vol. 55,
Ph.D. degrees from the University of Twente,
no. 11, pp. 3159–3166, Nov. 2008.
Enschede, The Netherlands, in 1974 and 1978,
[22] B. Kaleli, T. van Hemert, R. J. E. Hueting, and R. A. M. Wolters,
respectively.
“Strain characterization of FinFETs using Raman spectroscopy,” Thin
He has been with Philips Research and NXP
Solid Films, vol. 541, pp. 57–61, Aug. 2013.
Research, Eindhoven, The Netherlands, since 1978.
[23] T. Rudenko, V. Kilchytska, N. Collaert, M. Jurczak, A. Nazarov, and
Since 2004, he has been a part-time Professor with
D. Flandre, “Carrier mobility in undoped triple-gate FinFET struc-
the MESA+ Institute for Nanotechnology and the
tures and limitations of its description in terms of top and sidewall
Chair of Semiconductor Components, University of
channel mobilities,” IEEE Trans. Electron Devices, vol. 55, no. 12,
Twente.
pp. 3532–3541, Dec. 2008.

You might also like