throbber
Acta Materialia 50 (2002) 3579–3595
`
`www.actamat-journals.com
`
`A reaction-layer mechanism for the delayed failure of
`micron-scale polycrystalline silicon structural films subjected
`to high-cycle fatigue loading
`C.L. Muhlstein a, E.A. Stach b, R.O. Ritchie a,∗
`
`a Materials Sciences Division, Lawrence Berkeley National Laboratory, and Department of Materials Science and Engineering,
`University of California, Berkeley, CA 94720-1760, USA
`b National Center for Electron Microscopy, Lawrence Berkeley National Laboratory, Berkeley, CA 94720-1760, USA
`
`Received 5 November 2001; received in revised form 18 March 2002; accepted 18 March 2002
`
`Abstract
`
`A study has been made to discern the mechanisms for the delayed failure of 2-µm thick structural films of n+-type,
`polycrystalline silicon under high-cycle fatigue loading conditions. Such polycrystalline silicon films are used in small-
`scale structural applications including microelectromechanical systems (MEMS) and are known to display ‘metal-like’
`stress-life (S/N) fatigue behavior in room temperature air environments. Previously, fatigue lives in excess of 1011
`cycles have been observed at high frequency (~40 kHz), fully-reversed stress amplitudes as low as half the fracture
`strength using a surface micromachined, resonant-loaded, fatigue characterization structure. In this work the accumu-
`lation of fatigue-induced oxidation and cracking of the native SiO2 of the polycrystalline silicon was established using
`transmission electron and infrared microscopy and correlated with experimentally observed changes in specimen com-
`pliance using numerical models. These results were used to establish that the mechanism of the apparent fatigue failure
`of thin-film silicon involves sequential oxidation and environmentally-assisted crack growth solely within the native
`SiO2 layer. This ‘reaction-layer fatigue’ mechanism is only significant in thin films where the critical crack size for
`catastrophic failure can be reached by a crack growing within the oxide layer. It is shown that the susceptibility of
`thin-film silicon to such failures can be suppressed by the use of alkene-based monolayer coatings that prevent the
`formation of the native oxide.  2002 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
`
`Keywords: Silicon; Fatigue; Thin films; MEMS; Self-assembled monolayer coatings
`
`1. Introduction
`
`The promise of revolutionary commercial products
`at small dimensions has fueled the rapid development
`
`∗ Corresponding author: Tel.: +1-510-486-5798; fax: +1-
`510-486-4881.
`E-mail addresses: cmuhlstn@uclink4.berkeley.edu (C.L.
`Muhlstein); roritchie@lbl.gov (R.O. Ritchie).
`
`of microelectromechanical systems (MEMS) and the
`enabling technologies of surface micromachining. Sili-
`con-based structural films have emerged as the domi-
`nant material system for MEMS because the microma-
`chining technologies for silicon are readily adapted
`from the microelectronics industry, and are compatible
`with fabrication strategies for the integrated circuits
`necessary for actuation and control of the systems.
`
`1359-6454/02/$22.00  2002 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
`PII: S 13 59 - 6 4 5 4 ( 02 )0 0 1 5 8 - 1
`
`GE-1028.001
`
`

`

`3580
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`these
`long-term durability of
`the
`However,
`microsystems may be compromised by the suscep-
`tibility of thin-film silicon to delayed failure during
`cyclic loading conditions in ambient air [1–8].
`Cyclic fatigue is the most commonly encoun-
`tered mode of failure in structural materials, occur-
`ring in both ductile (metallic) and brittle (ceramic)
`solids
`(although
`the mechanisms
`are
`quite
`different) [9]. The mechanistic understanding of
`fatigue together with the use of damage/fracture
`mechanics to describe its effect at continuum
`dimensions has allowed for the reliable design and
`operation of innumerable macro-scale structures,
`such as aircraft airframes and engines. At
`the
`micro-scale, the fatigue of ductile materials is attri-
`buted to cyclic plasticity involving dislocation
`motion that
`causes
`alternating blunting and
`resharpening of a pre-existing crack tip as it
`advances
`[10].
`In contrast, brittle materials
`invariably lack dislocation mobility at ambient
`temperatures, such that fatigue occurs by cycle-
`dependent degradation of the (extrinsic) toughness
`of the material in the wake of the crack tip that
`developed from preexisting material inhomogen-
`eities [11]. Prior to this work, the relevance of these
`fatigue mechanisms to silicon films had yet to be
`established.
`Silicon is generally regarded as a prototypical
`brittle material; dislocation activity is generally not
`observed at low homologous temperatures (below
`~500 °C) and there is little evidence of extrinsic
`toughening, such as grain bridging or microcrack-
`ing [12]. Moreover, silicon is not susceptible to
`environmentally-induced cracking (i.e., stress-cor-
`rosion cracking) in moist air or water [13–15] at
`growth rates measurable in bulk specimens. These
`observations strongly suggest that silicon should
`not fatigue at room temperature. Indeed, there has
`been no evidence to date that bulk silicon is sus-
`ceptible to fatigue failure. However, there is sub-
`stantial evidence that cyclically-stressed, micron-
`scale, silicon films can fail prematurely under high-
`cycle fatigue loading [1–8,16].
`The observation that silicon thin films can fail
`under cyclic loading was first reported by Connally
`and Brown a decade ago [1]. Since then,
`the
`present authors and others [2–7] have confirmed
`that 2 to 20 µm thick single crystal and polycrystal-
`
`line silicon films can fail in fatigue at stresses as
`low as half their (single-cycle) fracture strength
`after more than ~1011 cycles. Despite such results,
`the mechanistic origins of why thin-film silicon
`should apparently suffer
`fatigue failure have
`remained elusive. Early studies highlighted the
`importance of water vapor and speculated that the
`mechanism may be associated with static fatigue
`of the native silica layer [1,7]. Other proposed
`explanations have involved dislocation activity in
`compression-loaded silicon (e.g.,
`[17]), stress-
`induced phase transformations [4], and impurity
`effects [4], although in no instance has conclusive
`experimental evidence been presented to support
`any of these mechanisms. Moreover, until now
`there has never been any direct observation of
`fatigue damage in micron-scale silicon, nor indi-
`cations on how it accumulates.
`A recent study by the authors [18], however,
`provided the initial experimental evidence that the
`crack initiation and growth processes involved in
`the apparent fatigue of silicon are confined to the
`amorphous SiO2 reaction layer that forms on sur-
`faces upon their exposure to air. In this paper, we
`present a mechanism for the apparent fatigue of
`silicon, termed reaction-layer fatigue, on the basis
`of prior stress-life fatigue data, a compliance tech-
`nique for monitoring the damage accumulation,
`and microstructural analysis using high-voltage
`transmission electron microscopy. Additionally,
`we suggest a method for suppressing the cyclic
`fatigue of silicon films through the use of alkene-
`based monolayer coatings that is validated with
`stress-life fatigue data.
`
`2. Experimental procedures
`
`The 2-µm thick silicon films were fabricated
`from the first structural polycrystalline silicon layer
`on run 18 of the MCNC/Cronos MUMPs pro-
`cess. This surface micromachining process utilizes
`low-pressure chemical vapor deposition (LPCVD)
`to manufacture n+-type (resistivity, r=1.9 × 10⫺3
`⍀·cm) polycrystalline silicon [19]. Wafer curvature
`measurements showed the film to have a compress-
`ive residual stress of about 9 MPa [19]; out-of-
`plane deformation due to a through-thickness
`
`GE-1028.002
`
`

`

`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`3581
`
`residual stress gradient could not be detected using
`white-light
`interferometry. Secondary ion mass
`spectroscopy (SIMS), referenced to known stan-
`dards, was used to quantify the concentration of
`hydrogen,
`carbon, oxygen,
`and phosphorous
`present.
`The elastic properties of polycrystalline silicon
`thin films approach the average behavior of ideal-
`ized polycrystalline materials. An average of the
`Voigt and Reuss bounds for a random, polycrystal-
`line aggregate (Young’s modulus, E=163 GPa,
`Poisson’s ratio, n=0.23 [20]) were used to estimate
`the elastic behavior of the material. The fracture
`strength of polycrystalline silicon typically ranges
`from 3 to 5 GPa depending on loading condition,
`specimen size, and test
`technique. The fracture
`toughness, Kc, is ~1 MPa√m [12,21].
`The microstructure of the films was charac-
`terized using transmission electron microscopy
`(TEM). Cross-sectional TEM specimens were pre-
`pared from the patterned films using standard lab-
`oratory practices [22]. Pairs of patterned chips con-
`taining the surface micromachined structures were
`glued together face-to-face, mechanically thinned,
`dimpled, and ion milled to the desired electron
`transparency. Diffraction contrast and high-resol-
`ution microscopy of the these specimens was per-
`formed using the Berkeley JEOL Atomic Resol-
`ution Microscope (ARM) at an operating voltage
`of 800 kV and a JEOL 3010 TEM operating at 300
`kV. Analytical characterization was accomplished
`using a Philips CM200 Field Emission microscope
`equipped with a Link Energy Dispersive Spec-
`trometer (EDS) and a Gatan Image Filter for elec-
`tron energy loss spectroscopy (EELS) and energy
`filtered imaging (EFTEM). Plan view observations
`of the grain morphology and oxide structure were
`accomplished using both the ARM and the Kratos
`High Voltage Electron Microscope (HVTEM)
`operating at 0.8–1.0 MeV. Plan view samples were
`prepared by simply lifting the micromachined
`structures off of the substrate using a tungsten
`probe tip and placing them onto 100 mesh clam
`shell grids. For HVTEM studies, no additional
`thinning was necessary to image through the entire
`2 µm thick samples.
`The stress-life (S/N) fatigue behavior of the
`polycrystalline silicon films was determined using
`
`a ~300-µm square, ~2-µm thick, surface microma-
`chined
`fatigue
`characterization
`structure,
`as
`described in ref. [5] (Fig. 1). Briefly, the notched
`cantilever beam specimen (~40-µm long, 19.5-µm
`wide, with a 13-µm deep, ~1 µm root radius notch)
`is attached to a large, perforated, plate-shaped mass
`and is electrostatically forced to resonate. On
`opposite sides of the resonant mass are interdigi-
`tated ‘fingers’ commonly known as ‘comb drives’;
`one side is for electrostatic actuation, the other pro-
`vides capacitive sensing of motion. The specimen
`is attached to an electrical ground, and a sinusoidal
`voltage (with no direct-current (DC) offset) at half
`the natural frequency is applied to one comb drive,
`thereby inducing a resonant response in the plane
`of
`the figure. These conditions generate fully
`reversed, constant amplitude, sinusoidal stresses at
`the notch, i.e., a load ratio (ratio of minimum to
`maximum load) of R=⫺1, that are controlled to
`better than 1% precision with a resolution of ~5%.
`Specimens were cycled to failure at resonance (~40
`kHz) in ambient air (~25 °C, 30–50% relative
`humidity) at stress amplitudes ranging from ~2 to
`4 GPa using the control scheme described in
`refs. [4,5].
`Specimens were prepared by removing the sacri-
`ficial oxide layer in 49% aqueous hydrofluoric acid
`(HF) for 21 / 2 or 3 min, drying at 110 °C in air,
`and subsequently mounting in ceramic electronic
`packages for testing [5]. In an attempt to suppress
`formation of the native oxide and access of moist-
`ure to the silicon surface, specific specimens were
`coated with an alkene-based monolayer of 1-octa-
`decene, C16H33CH=CH2, after removal of the sacri-
`ficial oxide. The monolayer was then applied to
`the surface of the silicon in a reactor containing
`a solution of one part 1-octadecene in nine parts
`hexadecane [23]. This hydrophobic monolayer
`bonds directly to the hydrogen-terminated surface
`atoms of the silicon film created by the exposure
`to hydrofluoric acid such that no oxide can form;
`it acts as an effective barrier to both oxygen and
`water [23].
`the
`The experimentally-measured motion of
`resonating fatigue characterization structure was
`used to determine the applied stress amplitude,
`natural frequency, and to monitor the accumulation
`of fatigue damage prior to failure. The magnitude
`
`GE-1028.003
`
`

`

`3582
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`Fig. 1. Scanning electron micrograph of the fatigue life characterization structure and notched cantilever beam specimen used in
`this investigation. The (a) mass, (b) comb drive actuator, (c) capacitive displacement sensor, and (d) notched cantilever beam specimen
`(inset) are shown.
`
`of the displacements was carefully calibrated, as
`detailed in ref. [5]. Finite element models were
`used to establish the relationship between the dis-
`placements and the maximum principal stress at
`the notch. It was previously demonstrated that
`measured changes in natural frequency may be
`attributed to damage accumulation in the specimen
`[7,24]. Thus, additional numerical models of struc-
`tures containing cracks were used to determine the
`relationship between crack length and natural
`frequency (i.e., compliance) and the stress-inten-
`sity factor, K. The models were constructed using
`a commercial software package (ANSYS v. 5.7);
`full details are reported in ref. [25]. In the present
`paper, such methods were used to measure in situ
`the propagation of nanometer-scale cracks by
`monitoring the change in natural frequency of the
`sample. Crack-growth rates were determined using
`a modified secant method applied over ranges of
`crack extension of 2 nm with a 50% overlap with
`the previous calculation window;
`the average
`crack-growth rate was calculated based on a linear
`fit of the experimental data. The maximum stress
`intensity immediately prior to failure was taken as
`an estimate of
`the fracture toughness of
`the
`material.
`After testing, the crack path and fracture sur-
`faces of the specimens were characterized using
`
`and
`(SEM)
`electron microscopy
`scanning
`HVTEM. To avoid corrupting any microstructural
`or fractographic features, neither SEM conductive
`coatings nor TEM thinning processes were used.
`To evaluate the possibility of specimen heating
`due to the large amplitude, high frequency stresses
`and the induced electrical current used to measure
`motion of the structure, high-resolution infrared
`(IR) imaging of the fatigue characterization struc-
`ture was performed in order to map temperature
`changes during testing. Thermal images were gen-
`erated by plotting the difference between IR
`images (12-bit
`resolution) collected while the
`structure was resonated at a constant stress ampli-
`tude, and at rest. Individual IR images were col-
`lected by averaging over 2 sec at an acquisition
`rate of 50 Hz. Temperature changes as small as 20
`mK could be detected with a spatial resolution of
`better than 8 µm.
`
`3. Results
`
`3.1. Microstructural analysis
`
`The microstructural analysis of the 2-µm thick
`polycrystalline silicon film, shown in the cross-sec-
`tional TEM image of Fig. 2a, revealed an equiaxed
`
`GE-1028.004
`
`

`

`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`3583
`
`(a) Microstructure of the polycrystalline silicon structural film showing a typical cross-sectional TEM image of the through-
`Fig. 2.
`thickness grain morphology. TEM images of defect types, showing: (b) 220 bright field image of the interior of the grain to highlight
`microtwins, stacking faults, and Lomer–Cottrell dislocation locks, (c) high resolution image of a Lomer–Cottrell lock (inset).
`
`grain morphology (grain size of ~100 nm), with
`no evidence of strong texture (from corresponding
`selected-area diffraction). No variations in micro-
`structure were apparent near features such as the
`root of the notch, as expected given the deposition
`and etching strategy used in the surface microma-
`chining process. The lack of a textured columnar
`structure, which is routinely observed in thin-film
`silicon [26], may be a result of the 900°C annealing
`used to dope the silicon with phosphorous and
`
`relax the residual stresses associated with growth
`of the film.
`the contaminants present
`SIMS analysis of
`revealed the interior of
`the film to contain
`苲 2 × 1018 atoms / cm3 hydrogen, 1 × 1018 atoms / cm3
`oxygen, and 6 × 1017 atoms / cm3 carbon [5], levels
`which are consistent with the processing history of
`the film. In addition, 1 × 1019 atoms / cm3 of phos-
`phorous were detected from the phosphosilicate
`glass used to dope the film. The films were found
`
`GE-1028.005
`
`

`

`3584
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`to be representative of materials used throughout
`micromachining and MEMS research and pro-
`duction. Oxygen concentrations
`in excess of
`苲 1018 atoms / cm3 at room temperature can be asso-
`ciated with precipitation of amorphous and crystal-
`line Si-O phases [27], although EELS of
`the
`interior of the grains revealed no segregation of
`oxygen or carbon in the film. Similarly, EFTEM
`imaging revealed no segregation of oxygen, car-
`bon, phosphorous, or nitrogen within the detect-
`ability limits of the technique. We conclude that
`that no precipitation of secondary species exists in
`the films.
`Diffraction contrast imaging was used to charac-
`terize the defect structure of the film and to confirm
`the absence of precipitation. A high magnification
`220 bright field image of the interior of a represen-
`tative grain, shown in Fig. 2b, reveals several dif-
`ferent types of lattice defects, including Lomer–
`Cottrell dislocation locks (a high resolution image
`of which is depicted in Fig. 2c), microtwins, and
`stacking faults. All dark areas of contrast in these
`cross sectional images, including polygonal fea-
`tures, were observed to correspond to one of these
`types of lattice defects and not to the presence of
`precipitates.
`
`3.2. Stress-life fatigue behavior
`
`Stress-life (S/N) data for the polycrystalline sili-
`con films from ref. [5], based on a total of 28
`fatigue specimens tested in room air, is shown in
`Fig. 3; fatigue lives, Nf, varied from ~10 sec to 34
`days (3 × 105 to 1.2 × 1011 cycles) for stress ampli-
`tudes ranging from ~2 to 4 GPa at R=⫺1. Two
`specimens were interrupted prior to failure for
`examination in the HVTEM. It is apparent that the
`polycrystalline silicon films display ‘metal-like’
`S/N behavior, with an endurance strength at 109–
`1010 cycles of roughly half the (single-cycle) frac-
`ture strength. Similar behavior has been seen in 20-
`µm thick films of single-crystal silicon cycled
`under similar conditions [4]. Stress-life fatigue
`tests were also conducted on specimens coated
`with the 1-octadecene monolayer. Thirteen speci-
`mens were tested to failure and five were interrup-
`ted prior to failure for examination in the HVTEM.
`Fatigue lives varied from ~7.5 sec to 25 days (3 ×
`
`Fig. 3. Typical stress-life (S/N) fatigue behavior of the 2 µm-
`thick, polycrystalline silicon at 苲40 kHz in moist room air under
`fully reversed, tension–compression loading [5].
`
`105 to 8.9 × 1010 cycles) for stress amplitudes rang-
`ing from ~1.4 to 3.3 GPa at R=⫺1 (Fig. 4a). In
`contrast to the specimens without the monolayer
`coating,
`the behavior
`is
`reminiscent of bulk
`brittle materials.
`
`3.3. Fractography and crack-path analysis
`
`SEM and TEM of both failed specimens and
`specimens interrupted during testing were used to
`evaluate fatigue damage, including the nature of
`the crack trajectory and the fracture surface mor-
`phology. Previous work [5] established that crack
`paths in polycrystalline silicon films during fatigue
`and subsequent overload failure are transgranular.
`Scanning electron microscopy at magnifications as
`high as 80,000× revealed a cleavage fracture mode
`with few distinctions in fracture surface mor-
`phology in the (presumed) fatigue and overload
`regimes.
`
`3.3.1. Overload fractures
`were
`surfaces
`Overload
`fracture
`(unambiguously) created by manually loading the
`fatigue test structure with a fixed (non-cyclic) dis-
`placement under an optical microscope; these con-
`ditions generate cracks that arrest due to the
`decreasing stress gradient associated with displace-
`
`GE-1028.006
`
`

`

`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`3585
`
`was clearly evident in the HVTEM. Examination
`of failed fatigue specimens and untested control
`samples revealed a stark difference in the native
`oxide found at
`the notch root. In the control
`samples, a native oxide of ~30 nm in thickness was
`uniformly distributed over the surfaces of the sam-
`ple, including the notch. 30-nm thick native oxide
`layers were also present on the tested fatigue sam-
`ple, except in the vicinity of the notch where the
`oxide layer was significantly thicker. Specifically,
`up to a three-fold increase in oxide thickness at the
`root of the notch was observed on samples that had
`been exposed to cyclic stresses (Fig. 6), a result
`confirmed on three different
`fatigued samples.
`Such enhanced notch-root oxidation was not
`observed in monotonically-loaded samples. Fur-
`thermore, no evidence of misfit dislocations due to
`a compressively-induced phase transformation was
`observed. As it is conceivable that the high loading
`frequency and induced electric currents may cause
`heating in the notch region, the role of thermal
`effects in the oxidation process was experimentally
`evaluated (Fig. 7), as discussed below.
`By interrupting fatigue specimens prior to fail-
`ure after testing at various stress amplitudes and
`examining them with HVTEM, several small
`cracks (on the order of tens of nanometers in
`length) were observed within the native oxide at
`the notch root (Fig. 8). The fact that these cracks
`were partially through the oxide layer indicates that
`they are stable cracks; indeed, this is the first evi-
`dence of stable fatigue cracking ostensibly in sili-
`con. Moreover, the size of the cracks was consist-
`ent with the compliance change of the sample
`predicted from the finite element modeling
`
`3.3.3. Infrared microscopy
`High-resolution (20 mK) infrared images were
`taken of the test structure during cyclic loading;
`results at rest and at increasing stress amplitude are
`shown in Fig. 7. Each image was created by plot-
`ting the difference between the IR image at rest
`and while resonating; the vertical scale represents
`temperature changes less than 1 K. The slight
`warming of the mass originates from friction of the
`silicon with air as the structure resonates at ~40
`kHz. However, the temperature of the structure
`does not rise significantly (⬍1 K) above ambient,
`
`Fig. 4. Comparison of the fatigue behavior of uncoated and
`monolayer-coated polycrystalline silicon thin films, showing (a)
`the respective S/N curves and (b) accumulated fatigue damage
`in terms of the change in natural frequency of the sample, fcrack,
`normalized by the natural frequency at the start of the test, f0.
`Note the reduced susceptibility of the coated polycrystalline
`silicon films to fatigue failure in (a), and the significantly
`enhanced life (by two orders of magnitude) compared to unco-
`ated polycrystalline silicon at comparable applied stress ampli-
`tudes in (b).
`
`ment-control in this configuration. Cracking in the
`silicon was confirmed to be transgranular cleavage
`(Fig. 5), with evidence of secondary cracking and
`microcracking consistent with the ‘slivers’ which
`appear on fracture surfaces (Fig. 5c).
`
`3.3.2. Fatigue fractures
`Although SEM studies were inconclusive in dis-
`cerning any differences in the fatigue and overload
`fractures, the distinct nature of these two processes
`
`GE-1028.007
`
`

`

`3586
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`Fig. 5. Fractography of failures in polycrystalline silicon films, showing (a) SEM and (b) HVTEM image of the crack trajectory
`out of the notch (in an unthinned specimen). (c) SEM images of the transgranular cleavage fracture surfaces of a long-life fatigue
`test (Nf=3.8×1010 cycles at σa=2.59 GPa). Horizontal arrow in (c) indicates the direction of crack propagation. Note the fine, needle-
`like features and debris on the fracture surface in (c).
`
`GE-1028.008
`
`

`

`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`3587
`
`Fig. 6. HVTEM image of the notch region in an unthinned polycrystalline silicon test sample, showing enhanced oxidation at the
`notch root that failed after 3.56×109 cycles at σa=2.26 GPa
`
`and there is no measurable change in the notched
`region. The absence of heating in the cantilever
`beam section clearly indicates that the enhanced
`notch-root oxidation is not thermally induced. As
`the notch root is (initially) the most highly stressed
`region of the structure, the process appears to be
`mechanical
`in origin. Furthermore, since oxide
`thickening is not observed under quasi-static load-
`ing,
`the phenomenon is associated with the
`cyclic loading.
`
`3.4. Fracture mechanics analysis
`
`The change in resonant frequency of the cantil-
`ever beam was monitored during each test to pro-
`vide a continuous measure of the specimen com-
`pliance. This frequency decreased monotonically
`(by as much as 50 Hz in the long-life tests) before
`eventual specimen failure at the notch (Fig. 9).
`This behavior suggests that the failure of the film
`occurs after progressive accumulation of damage,
`e.g., by the stable propagation of a crack. Indeed,
`
`the longer the life of the specimen, the larger the
`decrease in beam stiffness. A method for relating
`such changes in resonant frequency and damage
`processes was analyzed in refs. [5] and [25]; in the
`current work, we utilize this technique to monitor
`compliance changes quantitatively throughout the
`fatigue test.
`The frequency change, which from experimental
`observation was reasoned to be due to localized
`oxidation and cracking at the notch root, was ana-
`lyzed using plane-stress finite element modal
`analyses with ANSYS [5]. The model suggested
`that for ~1 nm of crack extension, a 1 Hz change
`in natural frequency should be observed. The cor-
`responding model for local notch root oxidation
`predicted an initial decrease in frequency with
`increasing oxide layer thickness (at a rate very
`similar to that induced by cracking) due to the rela-
`tively low elastic modulus of the SiO2; specifically,
`a 1 nm increase in oxide thickness resulted in a
`~0.5 Hz decrease in resonant
`frequency. The
`numerical models imply that the measured changes
`
`GE-1028.009
`
`

`

`3588
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`Fig. 7. High resolution (20 mK) infrared (IR) images of the fatigue characterization structure at rest and at increasing stress amplitude
`at ~40 kHz. Maximum temperature variations were less than 1 K above ambient and were not measurable in the notched cantilever
`beam specimen.
`
`in resonant frequency are consistent with processes
`occurring on length scales commensurate with the
`native oxide thickness. Accordingly, in the present
`work, estimates of the crack length were made dur-
`ing the tests and are plotted in Fig. 9; these esti-
`mates reveal crack sizes that were less than 50 nm
`throughout the entire fatigue test. As described pre-
`viously, HVTEM studies provided direct confir-
`mation for these damage processes.
`
`3.4.1. Crack-growth rates
`As noted above, in situ measurements of the
`change in natural frequency during the fatigue test
`were used to determine the crack length, a, as a
`function of time or cycles [25]. From such data,
`crack-growth rates, da/dN, were calculated over
`increments of 2 nm in crack extension; a represen-
`tative result in the form of da/dN as a function of
`a is shown in Fig. 10. Crack lengths during the
`
`GE-1028.010
`
`

`

`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`3589
`
`Fig. 8. HVTEM image showing stable cracks, ~50 nm in
`length, in the native oxide formed during cyclic loading of a
`notched, polycrystalline silicon beam. Testing of this sample
`was interrupted after N=3.56×109 cycles at a stress amplitude
`σa=2.51 GPa. Image was intentionally defocused to facilitate
`the observation of the cracks.
`
`Fig. 9. Representative damage accumulation in polycrystalline
`silicon, shown by experimentally measured decrease in resonant
`frequency, fcrack, with time during a fatigue test (Nf=2.23×1010
`cycles at σa=3.15 GPa) and the corresponding computed
`increase in crack length, a.
`
`Fig. 10. Representative computed fatigue-crack growth rates,
`da/dN, as a function of the crack length, a, during a fatigue test
`of a polycrystalline silicon thin film (Nf=2.32×1010 cycles at
`σa=3.15 GPa). Growth rates were determined from the change
`in specimen compliance monitored throughout the test. Note
`that the entire fatigue process of crack initiation and growth
`until the onset of catastrophic failure occurs for crack sizes
`below ~50 nm, i.e., within the native oxide layer.
`
`fatigue process can be seen to remain under ~50
`nm, consistent with the fatigue process occurring
`in the oxide layer. Growth rates are vanishingly
`small1 and progressively decrease with increasing
`crack length during the test.
`
`3.4.2. Fracture toughness and critical crack size
`If the crack-driving force at failure is taken as
`an estimate of the fracture toughness, the resistance
`of the material to unstable crack growth can be
`determined. The crack length estimates at failure
`were used with plane-strain models of the stress
`intensity for the notched cantilever beam structure;
`the relationship between crack length, applied
`forces, and the stress intensity are detailed in ref.
`[25]. Results
`from the polycrystalline silicon
`
`1 The absolute values of these crack-growth rates are clearly
`physically unrealistic. However, the growth rates reported are
`average values associated with the number of cycles at the mea-
`sured natural frequency (~40 kHz) for the crack to grow over
`2 nm increments, assuming that the crack is growing continu-
`ously. Although this is the standard way of computing fatigue-
`crack growth rates [28], in reality the crack may not propagate
`every cycle and the crack front may not advance uniformly.
`
`GE-1028.011
`
`

`

`3590
`
`C.L. Muhlstein et al. / Acta Materialia 50 (2002) 3579–3595
`
`fatigue tests (Fig. 11) give an average fracture
`toughness of ~0.85 MPa√m, consistent with that of
`the native oxide [29]. However, it is important to
`note the relationship between the critical crack size
`at final failure, ac, and the thickness, ho, of the SiO2
`reaction layer when K=Kc. Such critical crack sizes
`are estimated in Fig. 12 for the range of maximum
`principal stresses used in this investigation. It is
`apparent that for applied stresses of 2 to 4 GPa,
`which caused failure in the present films after 105
`to 1011 cycles, the critical crack sizes are less than
`50 nm,
`i.e., comparable to, or
`less than,
`the
`observed oxide layer thicknesses (i.e., acⱕho). This
`indicates that the entire process of fatigue-crack
`initiation, propagation, and the onset of cata-
`strophic (overload) failure all occur within the
`oxide layer.
`
`4. Discussion
`
`4.1. Reaction-layer fatigue
`
`The cyclic fatigue of brittle materials is usually
`associated with the degradation of
`(extrinsic)
`toughening mechanisms in the crack wake [9].
`Such toughening arises from crack-tip shielding,
`
`Fig. 11. Computed fracture toughness, Kc, values determined
`from the estimated crack length immediately prior to failure,
`i.e., from the critical crack size, ac, for the given applied stress
`amplitude. The average fracture toughness of 0.85 MPa√m is
`consistent with failure within the native oxide on polycrystalline
`silicon at room temperature.
`
`Fig. 12. Computed estimates of the critical crack size, ac, as
`a function of the applied stress amplitude, σa, in the polycrystal-
`line silicon fatigue characterization structure. Critical crack
`sizes for the stress amplitudes used in the present tests (σa~2
`to 4 GPa) are less than ~50 nm, indicating that the onset of
`final failure of the structure occurs at crack sizes within the
`native oxide reaction layer.
`
`ceramic
`(non-transforming)
`brittle
`in
`which
`materials generally results from mechanisms such
`as grain bridging. Under cyclic loading, frictional
`wear in the sliding grain boundaries can lead to a
`progressive decay in the bridging stresses (e.g.,
`refs. [9,30,31]). The fatigue of brittle materials is
`therefore invariably associated with intergranular
`failure. When such materials fail transgranularly,
`e.g., as in commercial SiC and sapphire [32,33],
`there is generally little or no susceptibility to
`fatigue failure. Since polycrystalline silicon always
`fails transgranularly [5,21,34], and there has been
`no reported evidence of extrinsic toughening
`mechanisms, the material would not be expected
`to be prone to cyclic fatigue. No precipitates were
`detected in the silicon films that may also enable
`conventional, brittle material fatig

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket