1994-Plate Impact Response of Ceramics and Glasses

download 1994-Plate Impact Response of Ceramics and Glasses

of 9

Transcript of 1994-Plate Impact Response of Ceramics and Glasses

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    1/9

    Plate impact response of ceramics and glasses

    G. F. Raiser , J. L. Wise, R. J. Clifton, D. E. Grady, and D. E. Cox 

    Citation: Journal of Applied Physics 75, 3862 (1994); doi: 10.1063/1.356066 

    View online: http://dx.doi.org/10.1063/1.356066 

    View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/75/8?ver=pdfcov 

    Published by the AIP Publishing 

    Articles you may be interested in 

    On the shattering of clusters by surface impact heating 

    J. Chem. Phys. 105, 8097 (1996); 10.1063/1.472663

    Effects of geometry in pressure–shear and normal plate impact recovery experiments: Three‐dimensional

    finite‐element simulation and experimental observation J. Appl. Phys. 80, 3267 (1996); 10.1063/1.363268

    Scaling of the fingering pattern of an impacting drop 

    Phys. Fluids 8, 1344 (1996); 10.1063/1.868941

    Transient acoustic near field in air generated by impacted plates 

    J. Acoust. Soc. Am. 99, 700 (1996); 10.1121/1.414646

    Modeling of Kipp‐Grady plate impact experiments on ceramics using the Rajendran‐Grove ceramic model 

     AIP Conf. Proc. 309, 749 (1994); 10.1063/1.46481

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

    http://scitation.aip.org/search?value1=G.+F.+Raiser&option1=authorhttp://scitation.aip.org/search?value1=J.+L.+Wise&option1=authorhttp://scitation.aip.org/search?value1=R.+J.+Clifton&option1=authorhttp://scitation.aip.org/search?value1=D.+E.+Grady&option1=authorhttp://scitation.aip.org/search?value1=D.+E.+Cox&option1=authorhttp://scitation.aip.org/content/aip/journal/jap?ver=pdfcovhttp://dx.doi.org/10.1063/1.356066http://scitation.aip.org/content/aip/journal/jap/75/8?ver=pdfcovhttp://scitation.aip.org/content/aip?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jcp/105/18/10.1063/1.472663?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jap/80/6/10.1063/1.363268?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jap/80/6/10.1063/1.363268?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pof2/8/6/10.1063/1.868941?ver=pdfcovhttp://scitation.aip.org/content/asa/journal/jasa/99/2/10.1121/1.414646?ver=pdfcovhttp://scitation.aip.org/content/aip/proceeding/aipcp/10.1063/1.46481?ver=pdfcovhttp://scitation.aip.org/content/aip/proceeding/aipcp/10.1063/1.46481?ver=pdfcovhttp://scitation.aip.org/content/asa/journal/jasa/99/2/10.1121/1.414646?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/pof2/8/6/10.1063/1.868941?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jap/80/6/10.1063/1.363268?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jap/80/6/10.1063/1.363268?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jcp/105/18/10.1063/1.472663?ver=pdfcovhttp://scitation.aip.org/content/aip?ver=pdfcovhttp://scitation.aip.org/content/aip/journal/jap/75/8?ver=pdfcovhttp://dx.doi.org/10.1063/1.356066http://scitation.aip.org/content/aip/journal/jap?ver=pdfcovhttp://scitation.aip.org/search?value1=D.+E.+Cox&option1=authorhttp://scitation.aip.org/search?value1=D.+E.+Grady&option1=authorhttp://scitation.aip.org/search?value1=R.+J.+Clifton&option1=authorhttp://scitation.aip.org/search?value1=J.+L.+Wise&option1=authorhttp://scitation.aip.org/search?value1=G.+F.+Raiser&option1=authorhttp://oasc12039.247realmedia.com/RealMedia/ads/click_lx.ads/www.aip.org/pt/adcenter/pdfcover_test/L-37/1060384706/x01/AIP-PT/JAP_ArticleDL_012716/AIP-APL_Photonics_Launch_1640x440_general_PDF_ad.jpg/434f71374e315a556e61414141774c75?xhttp://scitation.aip.org/content/aip/journal/jap?ver=pdfcov

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    2/9

    Plate impact response of ceramics and glasses

    G. F. Raisera)

    Division of Engineering, Brown University, Providence, Rhode Island 02912

    J. L. Wise

    Organization 1433, Sandia National Laboratories, Albuquerque, New Mexico 87185

    R. J. Clifton

    Division

    of

    Engineering, Brown Vniversity, Providence, Rhode Island 02912

    D. E. Grady and D. E. Cox

    Organization 1433, Sandia National Laboratories, Albuquerque, New Mexico 87185

    (Received 3 March 1993; accepted for publication 11 January 1994)

    Soft-recovery plate impact experiments have been conducted to study the evolution of damage

    in polycrystalline Al,O, samples. Examination of the recovered samples by means of scanning

    electron microscopy and transmission electron microscopy has revealed that microcracking

    occurs along grain boundaries; the cracks appear to emanate from grain-boundary triple points.

    Velocity-time profiles measured at the rear surface of the momentum trap indicate that the

    compressive pulse is not fully elastic even when the maximum amplitude of the pulse is

    significantly less than the Hugoniot elastic limit. Attempts to explain this seemingly anomalous

    behavior are summarized. Primary attention is given to the role of the intergranular glassy phase

    which arises from sintering aids and which is ultimately forced into the interfaces and voids

    between the ceramic grains. Experiments are reported on the effects of grain size and glass

    content on the resistance of the sample to damage during the initial compressive pulse. To

    further understand the role of the glass, plate impact experiments were conducted on glass with

    chemical composition comparable to that which is present in the ceramic. These experiments

    were designed to gain further insight into the possibility of “failure waves” in glasses under

    compressive loading.

    I. INTRODUCTlON

    The dynamic recovery plate impact experiment has

    been considered an attractive method for evaluating the

    brittle behavior of ceramics for two main reasons. First,

    since microdamage in ceramic materials takes place at very

    rapid rates, these experiments provide a means of initiating

    microcracks but removing the loads before the microcracks

    coalesce into large-scale cracks and rupture the material.

    Second, if the plate thicknesses and geometries are chosen

    correctly, it is possible to subject the central region of the

    specimen to a well-known stress history and still recover it

    for microscopy studies.

    Several studies have utilized some of these ideas in

    plate impact experiments designed to study damage mech-

    anisms in ceramics and ceramic-related materials. Yaziv’

    used a double-impact technique which led him to charac-

    terize tensile damage as comprising a “spa11 zone.” Longy

    and Cagnoux’ used spa11 and recovery experiments to

    study how certain microstructures of alumina ceramics af-

    fect their spa11properties and their Hugoniot elastic limit

    (HEL). Louro3 tested different aluminas under various

    stress pulse durations and magnitudes to highlight how

    porosity and grain size alter their dynamic damage prop-

    erties. Stress histories in the ceramic were unfortunately

    indeterminate, and post-test analyses were limited to gross

    effects on recovered fragments. This is because these inves-

    akurrent address: Washington State University, Pull man, WA.

    tigations were conducted at very large stresses and impact

    velocities, so reusable target holders could not be con-

    structed to stop projectiles at impact. Consequently, flyer

    plates were constantly reaccelerated by the projectile, giv-

    ing rise to multiple impacts on the target.

    Stopping the projectile can be accomplished if impact

    velocities are reduced. Although low velocities may bring

    stresses below the reported HEL values for these materials

    (4-9 GPa), there are reasons to expect that some inelastic

    response occurs below this “elastic threshold.“4 Measured

    HEL values are often higher than actual values because full

    decay of the precursor has not occurred at the distances

    where the HEL is measured, Also, the existence of residual

    stresses (as in the case of ceramics) can promote load-

    induced microcracking.5-7

    In fact, given that in a ceramic

    with medium to large grain size certain facets will micro-

    crack spontaneously during processing, the strength of a

    ceramic’s elastic response most likely varies continuously

    from low to high stresses due to the wide distribution of

    residual stresses present in the material. For these reasons,

    microdamage can be expected at stress levels below the

    HEL.

    Recovery experiments have been successfully

    conducted’>’ in an extensive study of Vistal, an CT-A1203

    from Coors Porcelain Company. Velocity-time and stress-

    time laser interferometry data, electron microscopy pio

    tures, and finite-element modeling were presented. Elastic,

    three-dimensional finite-element calculations were also car-

    ried out to show that the desired stress-history profile was

    3882 J. Appl. Phys. 75 (a), 15 April 1994 0021-8979/94/75(8)/3862/8/$6.00 @ 1994 American Institute of Physics

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    3/9

    imposed on the specimen.g Results of the Vista1 study

    pointed to the critical role played by the intergranular

    glassy phase in this material, which is present in all sin-

    tered Al,Os ceramics.

    A more complete understanding is currently needed

    for the compressive and tensile damage processes that oc-

    cur in aluminas below the HEL. In addition, an explana-

    tion is required for the influence of grain size and, espe-

    cially, the intergranular glassy phase on the damage

    resistance of these ceramics. To this end, a series of alumi-

    nas of different microstructures has been subjected to com-

    parable stress levels and stress histories. Velocity-time pro-

    files and electron micrographs are compared and

    conclusions are made regarding the importance of grain

    size and glassy phase in these materials.

    To understand further the role of the glass, experi-

    ments have been conducted on an aluminosilicate, i.e.,

    Corning’s Cl723 glass, with a chemical composition that is

    similar to the one found at triple points wi thin the Vista1

    alumina.8’10 These experiments are intended to aid in un-

    derstanding the recently reported phenomenon of “failure

    waves”

    in glass.1’-*3 According to Brar and

    co-workers,111’2 when soda lime glass is impacted above its

    HEL, a wave with speed 2.2*0.2 mm/ps [less than both

    the longitudinal and shear wave speeds) is propagated into

    the glass from the impact face. Behind this wave the spa11

    strength is reported to drop and the transverse stress is

    reported to increase. They interpret this behavior as a de-

    crease in shear strength and a comminution of the glass.

    Kane1 and co-workersI used thick flyer plates to impact

    K19 glass and interpreted a small step in rear-surface par-

    ticle velocity as evidence of a recompression from the failed

    material, and hence, evidence of a failure wave, Moreover,

    they reported that failure waves are observed near or below

    the HEL of K19 glass. Some questions arise from these

    conclusions. For example, how can extensive cracking oc-

    cur in a nonporous material under a state of uniaxial com-

    pression, where any crack would be opening against high

    compressive stresses which would tend to close it? Also, if

    this phenomenon is indeed a material property, then why

    does it start immediately at the surface, but not inside the

    material where the compressive stress behind the leading

    wave front is unattenuated with distance of propagation

    and is held for an extended time prior to failure? To un-

    derstand the nature of a failure wave, spa11experiments

    have been conducted on Cl723 aluminosilicate glass spec-

    imens with different initial surface roughnesses. These tests

    were intended to probe not only the spall strength of the

    glass after compression to different stress levels, but also to

    investigate the possibility that the failure wave phenome-

    non is a surface effect in the sense that it emanates from

    surface irregularities which cause nonplanar waves to de-

    velop.

    II. PROCEDURE

    A. Recovery experiments on A1203

    These experiments were conducted in the Plate Impact

    Facility at Brown University. A diagram of the essential

    FIG. 1. Star-flyer recovery experiment configuration.

    elements inside the target chamber is shown in Fig. 1. A

    projectile trips several velocity pins as it emerges from the

    63.5 mm .gun barrel and strikes the steel anvil, which is

    bolted down inside the target chamber. Dimensions of all

    critical parts are set such that the star flyer (Ti6Al4V)

    impacts the specimen and the projectile nosepiece (steel)

    hits the anvil simultaneously. This stopping of the projec-

    tile prevents additional impacts of the flyer on the speci-

    men. Moreover, since the flyer has a lower impedance than

    the specimen, there is contact only for a time equal to the

    longitudinal wave round-trip time in the flyer. The star

    essentially bounces oft the specimen. The initial compres-

    sive pulse in the momentum trap reflects from its. rear

    surface (becoming tensile) and returns to the weakly

    bonded specimen interface, causing the momentum trap

    (Hampden steel hardened to 60-62 Ro) to separate from

    the specimen and fly off with the momentum that the flyer

    lost. The star-shaped flyer minimizes the effects of lateral

    unloading waves in a central octagonal region.g*14A short

    tensile pulse arises in the specimen due to an intended gap

    at the specimen-momentum trap interface. Thi s gap can be

    increased by sputtering a thin (several micrometers) layer

    of material onto the four corners of the momentum trap

    prior to bonding it to the specimen. Moreover, it can be

    eliminated by careful lapping and assembly of the two

    plates. The compressive pulse will reflect as tension for the

    time required to close the gap (usually 5-60 ns). This

    tensile pulse reflects from the impact face as a compressive

    pulse and travels into the momentum trap prior to separa-

    tion. The specimen is left at rest inside the aluminum

    holder and taken out for microscopy analysis. The La-

    grangian t-X diagram in Fig. 2 shows how these waves

    traverse the plates and indicates the sequential order in

    which they arrive at the rear surface of the momentum

    trap. The laser interferometer system monitors these plane

    waves at this surface, at four separate points within the

    plate’s central octagonal region, giving redundant, nearly

    identical, velocity-time and derived stress-time data.

    Each plate is lapped and polished to measured flat-

    nesses within one wavelength of a monochromatic light

    J. Appl. Phys., Vol. 75, No. 8, 15 April 1994

    Raiser et

    al.

    3863

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    4/9

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    5/9

    NosepieceIlmpactor

    FIG. 4. Spa11 xperiment configuration.

    aluminum impactor plate mounted on its forward face, and

    the supporting nosepiece has a counterbore in the center,

    making the rear face of the impactor a free surface. Upon

    impact, compressive stress states are induced in the impac-

    tor and target plates. Wave reflections from free surfaces

    interact to form a tensile stress state inside the glass spec-

    imen. The location where this tension first occurs can be

    controlled by varying plate thicknesses. The Lagrangian

    t-X diagram of Fig. 5 maps the stress state histories at

    different locations wi thin the impacting plates. In region 2

    the stress state is compressive while in region 5 the stress

    state first becomes tensile in the glass specimen. The par-

    ticle velocities of states, 0, 3, and 6 are monitored using

    VISAR~ interferometry. If there is no failure wave and the

    magnitude of the tensile stress in region 5 is less than the

    spa11strength of the glass, then the particle velocity will

    drop nearly to zero in region 6. The dashed line (repre-

    senting a failure wave) marks the boundary between failed

    material (on the left-hand side) and unfailed material (on

    the’right-hand side). In this figure state 5 first occurs to the

    left-hand side of the failure wave. If the glass has failed in

    this region, it will have no tensile strength and a “spa11

    plane” will be created. Left-hand-going waves that reflect

    FIG. 5. Lagrangian t-X diagram for the spa11experiment.

    from this surface will be weak, and will return to the glass

    rear surface through region 6 causing little or no reduction

    in free surface particle velocity. The dotted line maps the

    path of a reported recompression wavei1-13 arising from

    the release wave interaction with the failure wave. The

    occurrence of such a wave is explained on the basis of the

    failed material having a lower impedance than the unfailed

    material.

    The 6061-T6 aluminum flyer plate is lapped and pol-

    ished on both sides to a roughness of 0.02 pm rms. The

    Corning aluminosilicate glass is cut, lapped, and polished

    by Precision Glass Products Co. of Oreland, PA. Its flat-

    ness is measured as in Sec. II A and is better than four

    rings. The 60-40 polish has an average roughness of 0.04

    pm rms. A thin ( -500 nm) coating of aluminum is ap-

    plied to the rear surface of the glass to give it the necessary

    reflectivity for laserinterferometry measurements. In shots

    requiring a “rough” glass impact face, the glass is lapped

    for 0.5 h using 15 pm B& powder, giving a surface rough-

    ness of 0.52 pm rms. An epoxy bond around the target-

    plate periphery holds the glass concentricially within a

    standard aluminum support ring. The impact surface of

    the glass is flush with the front face of this ring. Impact

    velocity and tilt are measured using coaxial shorting pins

    mounted in this ring. The particle velocity history is mea-

    sured using a VISAR.~’ The fringe constants are 0.4028

    mm/,&fringe for shots GLASS1 and GLASS2 and 0.1988

    mm/,&fringe for shots GLASS3 and GLASS4. The two

    quadrature records and the beam intensity variation are

    recorded on LeCroy digitizers (previously described).

    Data reduction is carried out using VISARSS, a program

    developed at Sandia.2*

    Ill. RESULTS AND DISCUSSION

    A. Recovery experiments

    To check the experimental approach, two shots were

    conducted using high-strength metal specimens that re-

    mained elastic throughout the loading history. Figure 6

    shows the data recorded from one of these shots (92-08))

    where the flyer is Ti6A14V (acoustic impedance pc, =27.7

    GPa ps/mm) and the specimen and momentum trap were

    Hampden steel (pcl=46.24 GPa ps/mm) hardened to 62

    Rc . The other elastic shot has the same features and there-

    fore is not presented. This figure and all subsequent figures

    display two velocity-time curves. One is the measured ex-

    perimental profile (solid line), and the other is the profile

    predicted for planar, elastic; longitudinal waves based on

    the experimental impact velocity (dashed line). Al though

    four points are monitored during all tests, no significant

    differences are observed between them,’ confirming the

    one-dimensional nature of the deformation in the central

    region of the plates. For this reason, and in the interest of

    graph clarity, only one representative data record is pre-

    sented for a given shot. As shown in shot 92-08, the in-

    tended stress history is imposed to reasonable accuracy.

    This observation confirms earlier three-dimensional finite

    element analyses by Espinosa et al9

    J. Appl. Phys., Vol. 75, No. 8, 15 April 1994

    Raiser et a/. 3865

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    6/9

    SHOT 92-08

    SHOT 89-05

    -.07 Exp. Data ElasticP-1

    1500

    0.06

    0.05

    0.04

    loo0

    0.03

    0.02 500

    0.01

    0.00250 0 260 500 750 1000 1250 iSo%

    Time [nsoc]

    SHOT 92-09

    0.06~.,....,....,....,....,....,...~

    - - - Elastic Prediction

    -

    t

    0.07

    0.06.05

    f 0.04.03

    i 0.02

    Time [nsoc]

    0.07

    0.06

    0.05

    .04

    0.03

    0.02

    0.01

    - - - Elastic Prediction

    1500 T

    %

    IWO

    1

    8

    500

    Tlmsme]

    SHOT 92-l 1

    m,....,--..,-...,-.-.,.~..,--.*,

    .Data --- ElasticF

    1 0 250 500 750 lool

    Time [nut]

    FIG. 6. Experimental particle velocity vs time data for shots 92-08 (elastic shot: steel specimen), 89-05 (Coors Vista1 specimen), 92-09 (Coors AD-995

    specimen), and 92-11 (Coors AD-999 specimen).

    Results from tests run on three different sintered alu-

    minas from Coors Porcelain Co. of Boulder, CO, are also

    presented in Fig. 6. Test parameters and ceramic material

    properties are summarized in Table I. All shots were con-

    ducted at nearly the same stress level, using the same star

    flyer and momentum trap materials (Ti6Al4V and Hamp-

    den steel, respectively). Comparisons between certain

    shots where specimens differ primarily in only one micro-

    structural feature allow conclusions to be made about that

    feature’s effect.

    Shots 89-05 and 92-09 were performed using Coors

    Vista1 and Coors AD-995, respectively. As shown in Table

    I, AD-995 has an average grain size similar to Vista1 (17

    pm for AD-995,20 pm for Vistal), but more intergranular

    TABLE I. Summa ry of recovery experiments on ceramics.

    glassy phase as indicated by the smaller preprocessing

    wt % AlLO,. As indicated in Fig. 6, there is an improve-

    ment in compressive response for AD-995 as less attenua-

    tion of the first pulse occurs. The additional glass in AD-

    995 appears to have reduced the intergranular residual

    stresses, requiring higher applied stresses to initiate dam-

    age. The lower residual stresses are attributed to the flow of

    glassy phase during processing to accommodate thermal-

    expansion anisotropy between grains.23 Shot 92-09 shows a

    large pull-back signal after the initial compressive pulse

    and severe attenuation of the middle (second) pulse, indi-

    cating that this material has experienced considerable dam-

    age in tension. A previously reported shot’ on Vistal, with

    a tensile pulse of comparable duration, shows a sustained

    Shot

    No.

    Impact

    stress

    WI%)

    Material

    Specific

    gravity

    Average

    grain size

    (pm)

    wt % Al,O,

    (preprocessing)

    Longitudinal

    wave speed

    (mm/w)

    Porosity

    (%b)

    89-05 1291 Coors Vista1

    3.99 20 99.9 10.8

    0.0

    92-09 1372 Coors AD-995

    3.89 17 99.5 10.45

    2.3

    92-l 1 1473 Coors AD-999

    3.96 3 99.9 10.2

    0.7

    92-14 1476 Browna HP Al,O,

    3.96 2.4 99.99 10.9

    0.8

    ‘Based on adaptation of process described by Staehler and co-workers (Ref. 22).

    866 J. Appl. Phys., Vol. 75, No. 8, 15 April 1994 Raiser

    et a/.

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    7/9

    SHOT 92-14

    XT

    0.07 15000.06

    z

    2 z

    i 0.05

    .a 0.04 1000

    0

    2 R

    8 0.03

    yg 0.02 500

    P

    (D

    E

    E

    tj 0.01

    z

    0.00250 0 250 500 750 1000 1250 150;

    Time [nsec]

    FIG. 7. Experimental particle velocity vs time data for shot 92-14 (hot-

    pressed alumina).

    tensile stress of approximately 400 MPa, indicating that

    Vista1 is superior to AD-995 in this regard. The increase in

    glassy phase appears to have made the AD-995 alumina

    weaker in tension by decreasing the tensile strength of the

    grain boundaries.

    Shot 92-11 was conducted on Coors AD-999. This ma-

    terial has nearly the same amount of glassy phase as Vista&

    but a much smaller average grain size (3 ,um for AD-999,

    20 pm for Vistal). Comparison of this shot with 89-05

    shows that the reduction in grain size results in a signifi-

    cant improvement in compressive behavior, as the first

    pulse is attenuated less than for any of the commercial

    aluminas tested. This behavior is consistent with the ex-

    pected reduction in residual stresses with decreasing grain

    size.‘$ The second pulse in these two shots reveals that the

    tensile strengthening is nearly insignificant, as expected for

    aluminas with similar amounts of glassy phase.

    These experiments demonstrate that the desired micro-

    structure for an alumina under dynamic loading at room

    temperature is one with a small grain size (improving com-

    pressive response) and a minimum of preprocessing addi-

    tives (improving tensile response). To test these conclu-

    sions, a high-purity, small-grain-sized alumina was

    processed in-house. The processing procedure was an ad-

    aptation of the procedure introduced by Staehler, Pre-

    debon, and Pletka.22 A vacuum hot press was performed

    on HPA-0.5AF 99.99%purity Al,O, with 50% 0.48 ,um

    particle distribution (obtained from Ceralox Corp. of Tuc-

    son, AZ) with the following conditions: 1400 “C, under

    axial pressure of 34.5 MPa for 2.5 h. The heating and

    cooling cycle temperatures were ramped at 5-10 “C/mm

    and pressure was maintained from 1400 “C until the end of

    the cooling cycle. The final density was measured as 3.952

    g/cm3 (99.2% of theoretical). Shot 92-14 (Fig. 7) is a

    successful experiment on this material. The compressive

    pulse was attenuated less than for any of the commercial

    ceramics tested. The grain size of this material was found

    to be 2.4 pm by the line intercept method on a typical SEM

    picture. The small grain size and reduced residual stresses

    are believed o be responsible or the strong compressive

    response. The tensile pulse rises nearly to its elastic level of

    1526 MPa. This marked improvement in tensile strength

    appears to result from the higher tensile strength of the

    grain boundaries. More grains are bonded directly to each

    other, eliminating many of the facets with weak, glassy-

    phase interfaces.

    B. Microscope examination

    Representative SEM micrographs made from cross

    sections of the central octagonal region of the recovered

    samples are presented in Fig. 8. The impact direction in all

    cases is left- to right-hand side. Figures 8 (a) and 8 (b)

    show typical compression-dominated and tension-

    dominated regions of the Coors AD-995 specimen used in

    shot 92-09. This alumina has a grain size similar to that of

    Vistal, but a lower preprocessing wt % of Al203 and con-

    sequently more intergranular glass. Compared with Vistal,

    this specimen was slightly better in compression, but much

    weaker in tension. The micrographs bear this out as the

    compression-dominated region [Fig. 8 (a)] shows some

    triple-point porosity, but relatively little microcracking,

    whereas the tension-dominated region [Fig. 8 (b)] shows a

    clear spallation with a large intergranular crack and addi-

    tional microcracking in the surrounding area. Figures 8 (c)

    and S(d) show typical compression-dominated and

    tension-dominated regions for shot 93-01 in which the

    specimen was hot pressed from Al,O, powder wi thout the

    addition of glass. The lack of visible microcracks in either

    the compression-dominated or tension-dominated regions

    is consistent with the essentially elastic response recorded

    for this specimen.

    C. Spall experiments

    The spa11experiments are summarized in Table II and

    experimental data is presented in Fig. 9. GLASS1 and

    GLASS2 were conducted at compressive shock stresses in

    the range 7.5-7.9 GPa, while compressive stresses for

    GLASS3 and GLASS4 were in the range 3.3-3.5 GPa. The

    two higher-amplitude curves in Fig. 9 are from GLASS1

    and GLASS2 and they show no unloading at the expected

    time ( - 1.9-2.0 ys). This is a clear indication that the

    glass has lost its spa11 trength in the region where tension

    first develops. The two lower-amplitude curves in Fig. 9

    are from GLASS3 and GLASS4, and those curves here

    show an almost complete drop in particle velocity at the

    estimated unloading time. This result shows that in these

    shots (3 and 4) the glass has a spa11 trength of at least 3.5

    GPa.

    These shots indicate that when the glass is subjected to

    a sufficiently large compressive stress it loses its spa11

    strength in the region where the reported failure wave has

    already passed. If the glass undergoes a sufficiently small

    initial compressive stress then the spa11 trength is retained.

    These observations are consistent with the published re-

    sults of previous failure wave experiments by Brar and

    co-workers11f’2 where this loss in spa11 trength is reported

    at elevated initial shock stress levels.

    J. Appl. Phys., Vol. 75, No. 8, 15 April 1994

    Raiser et a/. 3867

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to

    103.37.201.8 On: Mon, 01 Feb 2016 13:59:41

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    8/9

  • 8/17/2019 1994-Plate Impact Response of Ceramics and Glasses

    9/9

    mens of different impact surface roughnesses, no evidence

    of a surface effect on failure waves has been obtained.

    IV. CONCLUSIONS

    By adhering to the requirements to achieve a reliable

    experimental design for star flyer recovery experiments

    (explained in previous studies by Kumar and Clifton,14

    and Espinosa et al. 9), and proven in the elastic shots of this

    investigation, the identification of dominant microstruc-

    tural failure mechanisms in A&O, specimens has been ac-

    complished. The experiment has made it possible to isolate

    and discuss compression-dominated damage processes and

    improvements as well as tensile damage processes and im-

    provements. In compression, grain size reduction lowers

    average residual stresses at triple junctions and grain

    boundaries and makes the material less susceptible to

    triple-point and intergranular flow, sliding, and micro-

    cracking. Reducing the amount of glassy phase (corre-

    sponding to a higher percentage of pure preprocessing

    powder) makes tensile damage less likely by improving

    grain-boundary strength. These conclusions were tested on

    a high-purity, small-grain-sized alumina, processed

    through hot pressing. The dynamic compressive and tensile

    properties were found to be superior to those of all other

    specimens tested.

    Future experiments on ceramics will focus on the pro-

    cessing and testing of an alumina with an ultrafine grain

    size. Recent results in lowering the costs and improving the

    efficiency of preparing high-purity, nanometer-sized pow-

    ders has brought attention to the possibility of large-scale

    production of “nanophase” materials.25’26 Current reports

    on other types of nanophase ceramics indicate that these

    materials have potential for easy, near-net-shape forming

    and grain bonding without the addition of sintering

    aids.2’*28 The development of a nanophase alumina is ex-

    pected to exhibit the best possible properties of this ce-

    ramic.

    The spa11 xperiments on an aluminosilicate glass sup-

    port the evidence of some shock-stress-dependent failure

    phenomenon. At high enough compressive stresses (e.g.,

    7.5-7.9 GPa for these tests) the glass loses its spa11

    strength in a region where the postulated failure wave has

    already passed. When the stress is reduced to 3.3-3.5 GPa,

    there i s a retention of spa11 trength in the same area. Ev-

    idence of a recompression, indicating impedance mismatch

    across the failure wave boundary, is uncertain-a small

    increase in free-surface velocity appears to occur, but at a

    later time than expected. Finally, there is no evidence to

    suggest that a failure wave arises from impact surface ir-

    regularities of the glass.

    ACKNOWLEDGMENTS

    For the ceramic work, the authors would like to thank

    Professor S. Suresh and Professor B. Sheldon for helpful

    discussions on damage mechanisms and processing details

    related to ceramics. In addition, thanks are due to C. Bull,

    M. Mello, B. Dean, H. Stanton, T. Kirst, and K. Markert.

    Free samples of Al,O, were provided by D. Ranney of

    Coors Porcelain Co. and complimentary alumina powder

    was supplied by D. Bussell of Ceralox Corp. This ceramic

    work and Brown University’s contribution to the glass re-

    search were supported by the NSF-MRG at Brown Uni-

    versity entitled “Micro-Mechanics of Failure-Resistant

    Materials.” For the glass research, P. L. Stanton and J. R.

    Asay are gratefully acknowledged for their support of the

    experiments at Sandia National Laboratories. Also, thanks

    are due to J. H. Gieske, E. D. Apodaca, and J. A. Moya for

    materials characterization and plate preparation. Portions

    of this work performed at Sandia National Laboratories

    were supported by the U. S. Department of Energy under

    Contract No. DE-AC04-76DP00789.

    ‘D. Yaziv, Ph.D. thesis, University of Dayton, 1985.

    ‘F. Longy and J. Cagnoux, J. Am. Cer am. Sot. 72, 971 (1989).

    ‘L. H. L. Louro, Ph.D. thesis, New Mexico Institute of Mining and

    Technology, 1988.

    4H D. Espinosa, G. Raiser, R. J. Clifton, and M. Ortiz, J. Hard Mater.

    3;285 (1992).

    ‘Y. Fu and A. G. Evans, Acta. Metall. 33, 1515 ( 1985).

    6A. G. Evans, Acta Metall. 26, 1845 (1978).

    ‘S. Suresh and J. R. Brockenbrough, Acta Metall. 36, 1455 (1988).

    sG. Raiser, R. J. Clifton, and M. Ortiz, Mech. Mater. 10, 43 (1990).

    ‘H. D. Espinosa, G. Raiser, R. J. Clifton, and M. Or&, J. Appl. Phys.

    72, 3451 (1992).

    ‘OS..M. Wiederhorn, R. J. Hockey, R. F. Krause, Jr., and K. Jakus, J.

    Mater. Sci. 21, 810 (1986).

    “N. S. Brar, S. J. Bless, and 2. Rosenberg, Appl. Phys. Lett. 59, 3396

    (1991).

    ‘*N S. Brar, Z. Rosenberg, and S. J. Bless, J. Phys. (Paris) IV 1, 639

    (1991).

    “G I Kanel, S. V. Rasorenov, and V. E. Fortov, in Shock Compression

    .

    of Condensed Matter-1991, edited by S. C Schmidt, R. D. Dick, J. W.

    Forbes, and D. G. Tasker (North-Holland, Amsterdam, 1992), pp.

    45 l-454.

    14P Kumar and R. J. Clifton, J. Appl. Phys. 48, 4850 (1977).

    15P: Kumar and R. J. Clifton, J. Appl. Phys. 48, 1366 (1977).

    l6 G. R. Fowles, G. E. Duvall, J. R. Asay, P. Bellamy, F. Feist mann, D.

    Grady, T. Michaels, and R. Mitchell, Rev. Sci. Instrum. 41, 984

    (1970).

    “M. Mello, V. Prakash, and R. J. Clifton, in Shock Compression of Con-

    densed Mafter-1991, edited by S. C. Schmidt, R. D. Dick, J. W. Forbes,

    and D. G. Tasker (North-Holland, Amsterdam, 1992), pp. 763-766.

    ‘s W. Tong, Ph.D. thesis, Brown University, 1991.

    191. 0. Owate and R. Freer, J. Am. Ceram. Sot. 75, 1266 (1992).

    “L. M. Barker and R. E. Hollenbach, J. Appl. Phys. 43, 4669 (1972).

    “L. M. Barker, Sandia National Laboratories Report SAND88-2788,

    1988.

    r2J. Staehler, W. Predebon, and B. Pletka, in Proceedings of the 12th

    Army Symposi um on Solid Mechanics, edited by S. C. Chou, 1991, p.

    469.

    ‘3 J. S. Reed, Introduction to the Principles of Ceramic Processing (Wiley,

    New York, 1988), p. 469.

    24J. E. Blendell and R. L. Coble, J. Am. Ceram. Sot. 65, 174 (1982).

    *‘T. Beardsley, Sci. Am. 267, 114 (1992).

    r6J. Guo, J. Solid State Chem. 96, 108 (1992).

    “M. J. Mayo, R. W. Siegel, A. Narayanasamy, and W. D. Nix, J. Mater.

    Res. 5, 1073 (1990).

    *‘M J Mayo, R. W. Siegel, Y. X. Liao, and W. D. Nix, J. Mater. Res. 7,

    973 (1992).

    J. Appl. Phys., Vol. 75, No. 8, 15 April 1994

    Raiser et

    a/. 3869

    s article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to