Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND...

9
JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009 287 Comparison of Au and Au–Ni Alloys as Contact Materials for MEMS Switches Zhenyin Yang, Daniel J. Lichtenwalner, Arthur S. Morris, III, Senior Member, IEEE, Jacqueline Krim, and Angus I. Kingon, Member, IEEE Abstract—This paper reports on a comparison of gold and gold–nickel alloys as contact materials for microelectromechan- ical systems (MEMS) switches. Pure gold is commonly used as the contact material in low-force metal-contact MEMS switches. The top two failure mechanisms of these switches are wear and stiction, which may be related to the material softness and the relatively high surface adhesion, respectively. Alloying gold with another metal introduces new processing options to strengthen the material against wear and reduce surface adhesion. In this paper, the properties of Au–Ni alloys were investigated as the lower contact electrode was controlled by adjusting the nickel content and thermal processing conditions. A unique and efficient switching degradation test was conducted on the alloy samples, using pure gold upper microcontacts. Solid-solution Au–Ni sam- ples showed reduced wear rate but increased contact resistance, while two-phase Au–Ni (20 at.% Ni) showed a substantial im- provement of switching reliability with only a small increase of contact resistance. Discussion of the effects of phase separation, surface topography, hardness, and electrical resistivity on contact resistance and switch degradation is also included. [2008-0049] Index Terms—Alloys, Au–Ni alloys, electrical contacts, mi- croelectromechanical systems (MEMS), microswitch, radio- frequency (RF) MEMS. I. I NTRODUCTION R ADIO-FREQUENCY (RF) microelectromechanical sys- tems (MEMS) switches demonstrate great potential for a spectrum of microwave applications. Compared with con- ventional field-effect transistors and p-i-n diodes, RF MEMS switches offer much lower power consumption, much better isolation, and lower insertion loss. One major obstacle for large-volume commercial application of RF MEMS switches is reliability. A recent industrial survey has shown that stiction and wear are the top two failure modes for MEMS actuators [1]. Manuscript received March 2, 2008; revised September 3, 2008. First pub- lished February 2, 2009; current version published April 1, 2009. This work was supported by the Extreme Friction MURI program under AFOSR Grant FA9550-04-1-0381 entitled “Multifunctional Extreme Environment Surfaces: Nanotribology for Air and Space.” Subject Editor M. Mehregany. Z. Yang and D. J. Lichtenwalner are with the Department of Materials Science and Engineering, North Carolina State University, Raleigh, NC 27695 USA (e-mail: [email protected]; [email protected]). A. S. Morris, III is with wiSpry Inc., Irvine, CA 92618 USA (e-mail: [email protected]). J. Krim is with the Department of Physics, North Carolina State University, Raleigh, NC 27695 USA (e-mail: [email protected]). A. I. Kingon was with the Department of Materials Science and Engineering, North Carolina State University, Raleigh, NC 27695 USA. He is now with the Division of Engineering, Brown University, Providence, RI 02912 USA (e-mail: [email protected]). Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identifier 10.1109/JMEMS.2008.2010850 For metal-contact-type MEMS switches, pure gold is widely used as a contact material to achieve low contact resistance. Stiction and wear are prone to occur between two soft adhesive contact surfaces while switching. Wear may be described as contact deformation and/or material redistribution, ultimately causing contact surface roughening and affecting local contact force and contact resistance. Contact contamination has also been discussed as a degradation and failure mechanism [2], [3]. There have been some efforts in the device-design field to address this problem of Au-to-Au contact reliability [4], [5]. Notably, Oberhammer and Stemme [6] introduced a mechan- ically bistable electrostatic switching mechanism to achieve a small passive closing force and a large active opening force for soft metal contacts. On the materials’ side, a few alternative metals and alloys have also been investigated as contact ma- terials for metal-contact MEMS switches. McGruer et al. [7] showed that ruthenium (Ru), platinum (Pt), and rhodium (Rh) were susceptible to contamination and the contact resistance increased after a characteristic number of cycles, while gold alloys with a high gold percentage showed no contact resistance degradation under the same test conditions. Coutu et al. [8], [9] showed that alloying gold with a small amount of palladium (Pd) or Pt extended the microswitch lifetimes, with a small increase in contact resistance. However, at present, comprehen- sive investigations of the effects of surface topography, alloy composition, and material microstructure on contact resistance and lifetime performance are lacking. In order to build a solid material knowledge base for microswitch designers, correla- tions among material properties, contacting performance (such as contact resistance and lifetime), and failure modes (such as stiction and wear/material transfer) need to be built based on systematic experimental data for different materials. The research in this area has been slowed by the long time requirement for fabricating a microswitch with a particular new contact material. Typically, these microswitches are fabricated in Si foundries, and only a limited range of materials may enter the fabrication facility. Second, the fabrication process must be optimized for each material, and it may take many months to fabricate a set of switches to test a candidate contact material. Materials’ compatibility and process integration issues must be addressed in advance for every material to be tested. Some convenient contact wear test facilities have been used to research the characteristics of microswitches. Notably, Hyman and Mehregany [10] used a modified XY Z mi- crometer station to investigate the contact physics and test the current-carrying capability of MEMS switches. Schimkat [11] used an accurate loading system to survey materials for 1057-7157/$25.00 © 2009 IEEE Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Transcript of Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND...

Page 1: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009 287

Comparison of Au and Au–Ni Alloys as ContactMaterials for MEMS Switches

Zhenyin Yang, Daniel J. Lichtenwalner, Arthur S. Morris, III, Senior Member, IEEE,Jacqueline Krim, and Angus I. Kingon, Member, IEEE

Abstract—This paper reports on a comparison of gold andgold–nickel alloys as contact materials for microelectromechan-ical systems (MEMS) switches. Pure gold is commonly used asthe contact material in low-force metal-contact MEMS switches.The top two failure mechanisms of these switches are wear andstiction, which may be related to the material softness and therelatively high surface adhesion, respectively. Alloying gold withanother metal introduces new processing options to strengthenthe material against wear and reduce surface adhesion. In thispaper, the properties of Au–Ni alloys were investigated as thelower contact electrode was controlled by adjusting the nickelcontent and thermal processing conditions. A unique and efficientswitching degradation test was conducted on the alloy samples,using pure gold upper microcontacts. Solid-solution Au–Ni sam-ples showed reduced wear rate but increased contact resistance,while two-phase Au–Ni (20 at.% Ni) showed a substantial im-provement of switching reliability with only a small increase ofcontact resistance. Discussion of the effects of phase separation,surface topography, hardness, and electrical resistivity on contactresistance and switch degradation is also included. [2008-0049]

Index Terms—Alloys, Au–Ni alloys, electrical contacts, mi-croelectromechanical systems (MEMS), microswitch, radio-frequency (RF) MEMS.

I. INTRODUCTION

RADIO-FREQUENCY (RF) microelectromechanical sys-tems (MEMS) switches demonstrate great potential for

a spectrum of microwave applications. Compared with con-ventional field-effect transistors and p-i-n diodes, RF MEMSswitches offer much lower power consumption, much betterisolation, and lower insertion loss. One major obstacle forlarge-volume commercial application of RF MEMS switchesis reliability. A recent industrial survey has shown that stictionand wear are the top two failure modes for MEMS actuators [1].

Manuscript received March 2, 2008; revised September 3, 2008. First pub-lished February 2, 2009; current version published April 1, 2009. This workwas supported by the Extreme Friction MURI program under AFOSR GrantFA9550-04-1-0381 entitled “Multifunctional Extreme Environment Surfaces:Nanotribology for Air and Space.” Subject Editor M. Mehregany.

Z. Yang and D. J. Lichtenwalner are with the Department of MaterialsScience and Engineering, North Carolina State University, Raleigh, NC 27695USA (e-mail: [email protected]; [email protected]).

A. S. Morris, III is with wiSpry Inc., Irvine, CA 92618 USA (e-mail:[email protected]).

J. Krim is with the Department of Physics, North Carolina State University,Raleigh, NC 27695 USA (e-mail: [email protected]).

A. I. Kingon was with the Department of Materials Science and Engineering,North Carolina State University, Raleigh, NC 27695 USA. He is now withthe Division of Engineering, Brown University, Providence, RI 02912 USA(e-mail: [email protected]).

Color versions of one or more of the figures in this paper are available onlineat http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/JMEMS.2008.2010850

For metal-contact-type MEMS switches, pure gold is widelyused as a contact material to achieve low contact resistance.Stiction and wear are prone to occur between two soft adhesivecontact surfaces while switching. Wear may be described ascontact deformation and/or material redistribution, ultimatelycausing contact surface roughening and affecting local contactforce and contact resistance. Contact contamination has alsobeen discussed as a degradation and failure mechanism [2], [3].

There have been some efforts in the device-design field toaddress this problem of Au-to-Au contact reliability [4], [5].Notably, Oberhammer and Stemme [6] introduced a mechan-ically bistable electrostatic switching mechanism to achieve asmall passive closing force and a large active opening force forsoft metal contacts. On the materials’ side, a few alternativemetals and alloys have also been investigated as contact ma-terials for metal-contact MEMS switches. McGruer et al. [7]showed that ruthenium (Ru), platinum (Pt), and rhodium (Rh)were susceptible to contamination and the contact resistanceincreased after a characteristic number of cycles, while goldalloys with a high gold percentage showed no contact resistancedegradation under the same test conditions. Coutu et al. [8], [9]showed that alloying gold with a small amount of palladium(Pd) or Pt extended the microswitch lifetimes, with a smallincrease in contact resistance. However, at present, comprehen-sive investigations of the effects of surface topography, alloycomposition, and material microstructure on contact resistanceand lifetime performance are lacking. In order to build a solidmaterial knowledge base for microswitch designers, correla-tions among material properties, contacting performance (suchas contact resistance and lifetime), and failure modes (such asstiction and wear/material transfer) need to be built based onsystematic experimental data for different materials.

The research in this area has been slowed by the long timerequirement for fabricating a microswitch with a particular newcontact material. Typically, these microswitches are fabricatedin Si foundries, and only a limited range of materials may enterthe fabrication facility. Second, the fabrication process must beoptimized for each material, and it may take many months tofabricate a set of switches to test a candidate contact material.Materials’ compatibility and process integration issues must beaddressed in advance for every material to be tested.

Some convenient contact wear test facilities have beenused to research the characteristics of microswitches. Notably,Hyman and Mehregany [10] used a modified X–Y –Z mi-crometer station to investigate the contact physics and testthe current-carrying capability of MEMS switches. Schimkat[11] used an accurate loading system to survey materials for

1057-7157/$25.00 © 2009 IEEE

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 2: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

288 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009

Fig. 1. Au–Ni phase diagram, after Okamoto and Massalski [16].

microrelay actuators. A modified nanoindentor was also usedto study the hot-switching degradation of MEMS contacts [12].One concern about these tests is that they do not closelymimic real microswitches. In particular, the tests have difficultyduplicating the switch geometry (mechanical moving parts aretypically microscale cantilever structures with metal films beingpatterned on the surfaces) and the contact geometry (microcon-tacts); these are important issues because the contact geometryhas a strong influence on the stress and heat distribution acrossthe contacts and, thus, on the switch performance and failuremechanism [13], [14].

In this paper, these issues were addressed by comprehen-sively testing an alloy series of Au–Ni (with varied nickelcompositions) and by utilizing a newly developed switchingdegradation test facility closely mimicking an actual MEMSswitch operation [15].

Aside from the alloys mentioned in the earlier paragraphs,Au–Ni has materials’ characteristics showing it to be a poten-tially promising alternative to gold. It has been reported thatAu–Ni alloy contacts yield much lower adhesion than pure goldcontacts [9]. Low-force (as low as 0.6 mN) stable contact andreliable reopening were also achieved in the previous researchreport. However, in the previous study, only samples with onecomposition (5% Ni) were investigated, using bulky samples(rivets) as test components instead of a MEMS test structure.Therefore, we have undertaken a study of a wider range of alloycompositions, tested under a configuration typical of MEMSswitches. The Au–Ni phase diagram is shown in Fig. 1 [16].Note that both metals are Face Centered Cubic (FCC) structuresand exist as a two-phase mixture at a relatively low temperatureunder equilibrium conditions. However, a metastable single-phase alloy may also be produced under the low processingtemperatures utilized for the film deposition. Thus, a com-parison of the metastable solid solutions and the two-phasemixtures of the same overall composition can be undertakenso that both microstructure and composition effects can beexamined.

The switching degradation test facility used for this paperutilizes the upper cantilevers from commercial RF MEMSswitches and tests these against alternative bottom contactmaterials, using a modified atomic force microscope (AFM)as described in the following. The emphasis is placed on acomparison of the alloys and correlations between materialproperties, contact resistance, and contact degradation.

Fig. 2. Configuration for the switching tests.

II. EXPERIMENT

We have designed a unique setup for monitoring MEMSswitching behavior by integrating a real MEMS cantilever(from commercial wiSpry switches [17]) and separate thin-film bottom contact materials under test into an AFM [15].The test configuration is shown schematically in Fig. 2. Thecontact is made between the upper pure gold microcontactsand bottom candidate contact materials. The contact resistanceevolution and the switching motion are both monitored in situ.The test switch is cycled on/off using a piezoelectric actuator.The switching mechanics is designed in such a way that contactresistance failure is highly sensitive to the contact surfacedegradation. During test, the MEMS cantilever is intentionallyflexed in both “on” and “off” states to ensure a calibratedcontact force. This prebent geometry also ensures a nearlyvertical contact motion (less than 2.5◦ tilting) for the upper mi-crocontacts with respect to the bottom contact pads. The contactarea then is always located near the front edge of the micro-contacts due to this slight tilting. Vertical oscillation of thebottom contact pads alternates the contact points between thecantilever edge (insulating) and the microcontacts, with thesetwo states corresponding to the electrical “open” and “closed”states. We are aware that, in the commercial wiSpry switcheswith which we are making comparison, the ultimate rise in re-sistance appears to correlate with substantial deformation of theswitch contacts, consistent also with other reports [7], [13]. Wehave therefore made this test more sensitive to contact deforma-tion by limiting the vertical travel of the piezoelectric actuator.As deformation occurs, the ON-state contact is affected, andresistance rises markedly. As shown later in Section III, this willoccur after dramatically fewer cycles than in service, primarilybecause the limited travel will result in an “out-of-contact”condition after substantially less contact deformation than inthe case of the commercial switch. Thus, we can accelerateour investigation of contact degradation as a function of contactmaterial while using the commercial cantilever in a test that isstrongly representative of the actual operation.

Further conditions are briefly summarized in the following.

1) The test is typically undertaken under “hot-switching”conditions. This term is used by the electrical switchcommunity for the condition where a field is present

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 3: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289

when switch opening and closing takes place. The opencircuit voltage (Voc) is set at 1 V. The measurementcurrent is 1 mA, a value reached after a current transient.Transient current was limited by adding a parallel anda serial resistor to the noninductive switch circuit. Theparallel resistor controls the open circuit voltage; theserial resistor limits the transient current. Our resultssuggest that the transient accelerates switch degradation,and the transient was therefore limited and kept constantfor all the comparative tests.

2) The contact force is calibrated as ∼150 μN/contact,which is up to twice the contact force of the wiSpy’smicroswitches under service conditions. This increasedcontact force was adopted to increase the deformationand wear rates but was also carefully controlled for thecomparative evaluations of the contact materials. Sincethe tests were conducted in ambient air, a relativelylarge contact force also helps to penetrate any surfacecontamination film that may be present and ensures ametal–metal contact.

Therefore, under the test conditions utilized in this investi-gation, the effective lifetimes are substantially shortened [15],which enables differences between contact materials to be morereadily distinguished.

Pure Au and Au–Ni alloy (with up to 20 at.% Ni) bottomcontact samples were prepared using ion-beam sputter deposi-tion. Xenon gas was used in the ion source due to its relativelylarge mass, reducing energetic bombardment of the growingfilm. The base pressure was < 3 × 10−7 torr, and the depositionpressure was 2 × 10−4 torr. The accelerating voltage for thebeam is 600 V and for the beam current is 20 mA. The substrateis a silicon wafer coated with a 1000-Å-thick thermal oxide.A 5-nm chromium (Cr) layer and a 10-nm molybdenum (Mo)layer were deposited as an adhesion layer and a diffusion barrierlayer, respectively. A 250-nm film of pure Au or Au–Ni (withvarious nickel compositions) was then deposited.

The Au–Ni alloy was prepared by cosputtering from a rotat-ing gold target partially covered with a piece of the alloyingmetal (Ni). The atomic percentage was then determined bythe fractional target area the nickel occupies (in percent areaor, equivalently, the angle θ in comparison to 360◦) and itssputtering yield (Y ) compared with that of gold. The requiredangle θ was calculated, as follows:

at.% A =YA · θA

YA · θA + YB · θB(1)

⇒ θ =360◦

1 + 1−at.%Aat.%A

· YA

YB

(2)

where A and B are the different materials. The sputtering yieldsof Au and Ni under Xe ion bombardment were taken fromOhring [18]. The in situ rotation of the target ensures the uni-formity of the film composition. The films were patterned usingstandard photolithography to facilitate the four-point probemeasurement of the contact resistance during the switchinglifetime tests. Ion-beam etching thereafter was used to removemetal films from the uncovered area. N-methyl pyrrolidinone

Fig. 3. Lattice parameter versus nickel composition for films deposited atroom temperature and at 200 ◦C, compared with bulk solid-solution alloys(from [20]).

solution and deionized water were used, in sequence, to dis-solve the photoresist layer and clean the sample, respectively.

Electrical resistivity data were obtained on as-deposited thinfilms (before patterning) using the four-point probe technique(MAGNE-TRON Model-700). A four-point sheet resistancecorrection [19] was applied. Microhardness data were obtainedusing a Hysitron Triboscope nanoindentor (Berkovich indentertip-142.5◦ included angle). Surface topographic images weretaken on the as-deposited thin films using a CPR AFM. Re-lated software was used to obtain roughness statistics andheight-profile analysis. A BRUKER AXS X-ray diffractometerwas used to identify the sample structure. RIBER LAS-3000X-ray photoelectron spectroscopy (XPS) was used to detect thesurface chemical composition (< 10 nm). The takeoff anglewas ∼75◦ from the sample surface. A MAC-2 analyzer and MgKα (1253.6 eV) source were used in the experiment.

III. RESULTS AND DISCUSSION

Gold alloy lattice parameters obtained by X-ray diffraction(XRD) are shown in Fig. 3 for films of various Ni contents.As shown in Fig. 3, the lattice parameter (measured in thedirection normal to the wafer surface) is plotted versus theexpected Ni at.%, compared with values for bulk solid-solutionalloys (represented by diamonds along the trend line) [20]. Theroom-temperature-deposited thin films (Fig. 3, triangles) showa larger than expected lattice parameter, consistent with theeffects of compressive stress. A series of Au–Ni samples wasdeposited with 200 ◦C in situ heating. For these heated samples(Fig. 3, squares), the lattice parameter fits the bulk literaturedata quite well, indicating that the compressive stress can beminimized by moderate in situ heating. More importantly, wecan conclude from the XRD data that all as-deposited samplesyield the metastable solid-solution phase (no nickel peaks areobserved) and that our Ni content calculation is correct as well.

In order to obtain the stable two-phase mixture, rapid thermalannealing (RTA) was conducted on some samples. Sampleswere heated at 600 ◦C for 30 s in flowing N2 gas. XRD spectra,before and after RTA, are shown in Figs. 4 and 5, respectively.In the spectrum of the RTA sample, Au (111), Au (200),

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 4: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

290 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009

Fig. 4. XRD spectrum of the Au–Ni (20 at.% Ni) before RTA.

Fig. 5. XRD spectrum of the Au–Ni (20 at.% Ni) after RTA.

TABLE IMATERIAL PROPERTIES OF Au–Ni FILMS

Ni (111), and Ni (200) peaks can be clearly identified. Theprecipitation of a nickel-rich phase is confirmed (not presentbefore the RTA). Therefore, the effects of solid-solution harden-ing versus precipitation hardening can be compared. Note that,however, if nickel precipitates are present at the surface, theyare susceptible to oxidization in air.

The properties of the Au and Au–Ni alloys are summarizedin Table I. It can be seen that the measured electrical resistivityand microhardness increase with increasing Ni content in thesingle-phase metastable alloys. Annealing the Au–Ni alloy(20 at.% Ni) to form the two-phase mixture significantly lowersthe resistivity and hardness compared with the preannealedvalues.

Fig. 6. Calculated and measured contact resistance versus nickel composition.

The bulk resistivity is not the key parameter for the MEMSswitches; rather, it is the contact resistance. We would like togain some insight into the factors controlling this contact resis-tance. The relationship between force and contact resistance forgold-contact MEMS switches has been investigated by previousresearchers [10], [21], [22]. For a contact radius greater thanthe electron mean free path, the Maxwell spreading resistancemodel can be used to calculate the contact resistance. In ourtest, the nominal contact radius is much larger than the meanfree path in gold (∼36 nm); hence, the model should be valid.The contact resistance Rc is described by

Rc =ρ

2rc(3)

where ρ is the resistivity and rc is the contact radius [23]. Whiletwo different materials are in contact, the resistivity is estimatedas the average of the two. The contact radius can be obtainedusing an asperity deformation model [22]

rc =

√F

ξπH(4)

where F is the applied force, H is the hardness of the material,and ξ is the coefficient of the material deformation mode.Different modes are given as ξ < 0.3, elastic deformation;0.3 < ξ < 0.75, elastoplastic deformation; and 0.75 < ξ < 1,plastic deformation.

As described elsewhere, the contact force in our test is about150 μN/microcontact [15]. Plastic deformation is assumedbased on this force over the hundred-nanometer-scale contactarea. For the contact between different materials, which is thecase in our experiments, the averaged hardness is used forthe calculation. The electrical resistivity, microhardness, andthe measured and calculated contact resistance data (againstupper pure Au microcontacts) for the pure Au and Au–Nisamples are shown in Table I. The calculated and measuredcontact resistances are also shown in Fig. 6. As the electricalresistivity and microhardness increase with the Ni percentage,the model (calculated values) shows an increase of contact

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 5: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 291

Fig. 7. AFM images of pure Au sample. (a) Topographic. (b) Error.(c) Random line height profile.

Fig. 8. AFM images of Au–Ni (20 at.% Ni, solid solution, as deposited)sample. (a) Topographic. (b) Error. (c) Random line height profile.

resistance. The experimental contact resistance data roughlyfollow this trend. For pure Au samples, the experimental datafit the calculated contact resistance data quite well. However,for the Au–Ni metastable solid-solution series, the measuredcontact resistance is much higher than the calculated value.Two-phase Au–Ni samples (20 at.% Ni) showed a much lowercontact resistance compared with the solid-solution alloy at thesame composition. Furthermore, the model, in this case, againcorrectly predicted the measured contact resistance based on thephysical properties of the films and the contact force.

Based on the analysis made earlier, it might be reasonable tosuggest that, for the Au–Ni solid-solution series, the mismatchbetween the calculated and measured contact resistances is dueto the fact that the roughness effect is not included in the model.To gain insight into this issue, a detailed topographic analysiswas performed using AFM.

A series of AFM Au–Ni topographic images with grainheight analysis and rms surface roughness data is shown inFigs. 7–11. Topographic and error images of a pure gold

Fig. 9. AFM images of Au–Ni (20 at.% Ni, two phases, after RTA) sample.(a) Topographic. (b) Error. (c) Random line height profile.

Fig. 10. Maximum height difference in the random line scan versus Nicomposition.

Fig. 11. RMS roughness versus nickel composition.

sample, a solid-solution Au–Ni (20 at.% Ni) sample, and a two-phase Au–Ni (20 at.% Ni) sample are shown, respectively, inFigs. 7–9, where height-profile line scans across the sample

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 6: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

292 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009

Fig. 12. Microcontacting tests on pure gold and Au–Ni alloys.

surfaces are also shown (note that the height scale bar in Fig. 6is different from that in Figs. 7 and 8). The maximum heightdifference in the line scans and rms surface roughness data areshown in Figs. 10 and 11. The topographic analysis can besummarized as follows.

1) The metastable single-phase Au–Ni alloys show a markedtrend of increasing roughness with increasing Ni content.

2) The rapid thermally annealed two-phase Au–Ni alloy(20 at.% Ni) shows an entirely different microstructurewith a surprisingly reduced roughness compared withthe preannealed state. The mechanism for the surfacesmoothing is not yet determined.

3) The Au and the two-phase Au–Ni alloy (20 at.% Ni)samples show comparable values of roughness despitevery different microstructures and grain sizes. The similarvalues of contact resistance should also be noted.

Generally, it is assumed that rough surfaces yield lower con-tact areas than smooth surfaces. As a result, measured contactresistances for Au and two-phase Au–Ni samples fit quite wellwith calculated values, while solid-solution Au–Ni samplesshow a higher contact resistance than expected. This paper callsfor a more advanced contact resistance modeling, in which thetopology effect of the contact surfaces needs to be addressed.Recently, some progress has been made in this area [24].

It is an initially surprising result that the rapid annealing of ametastable solid-solution Au–Ni (20 at.%) sample resulted in amuch smoother surface for the two-phase structure. Neverthe-less, in the application of metal contacts for MEMS switches,this material processing method enables us to achieve a goodcombination of low surface roughness, low contact resistance,and increased material hardness.

Pure Au and Au–Ni thin-film test structures were subjectedto the switching degradation tests. Hot-switching conditionswere selected for the comparisons. The test conditions andunique electrical failure detection mechanism are as describedin Section II. The test results are shown in Fig. 12, showing theclosed-state resistance as a function of switching cycle (notethat a test cantilever has two microcontacts; thus, the resistancein Fig. 12 is roughly twice the contact resistance values in

Fig. 13. AFM images of the worn microcontacts on the Au cantilever andthe corresponding bottom electrodes in the lifetime tests of Au against Au andAu against solid-solution Au–Ni (20 at.% Ni). (a) Worn pure Au microcontacton the MEMS cantilever in Au-against-Au test. (b) Worn area on the pure Aubottom electrode corresponding to the microcontact shown in (a). (c) WornAu microcontact in Au-against-Au–Ni (20 at.% Ni) test. (d) Worn area onthe Au–Ni (20 at.% Ni) bottom electrode corresponding to the microcontactshown in (c).

Fig. 6). Failure is indicated by the contact resistance increasingtoward the open circuit value. It is clear that the number ofcycles before failure increases as the percentage of Ni in thealloy increases.

Typical worn microcontacts on the Au cantilever, as well asthe corresponding bottom electrodes in the degradation tests ofAu against Au and of Au against solid-solution Au–Ni (20 at.%Ni), are shown in Fig. 13. As indicated by the AFM pictures,the top electrode wear occurs predominantly on one edge, dueto a small angle between the top and the bottom contacts. Thebottom contacts show material that has been transferred fromthe upper dimple onto the lower contact. In the case of theAu–Ni alloy (which has seen more than twice as many cycles),this material is distributed over a slightly larger area than thecase of pure gold. However, Fig. 13 also shows (and moreconvincingly confirmed in AFM analysis, not included) that thedepression in the center of the contact area, presumably dueto plastic deformation, is substantially less in the case of theAu–Ni alloy.

Information revealed by the test can be summarized asfollows.

1) Low initial contact resistance was achieved for all thetests. A slight decrease of the initial contact resistancecan be observed after a number of cycles. Notably, thecontact resistance of solid-solution Au–Ni (20 at.% Ni)drops from the initial value of ∼3.5 Ω to the stable valueof ∼1.8 Ω after a couple hundred switching cycles. This

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 7: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 293

“burn-in” period is less obvious for all the other sam-ples. The low initial contact resistance suggests that anysurface contamination film is insignificant. The relativelylarge contact force (150 μN/microcontact) utilized in thetests, as well as the possible arc energy dissipation andtransient current heating, may remove surface contam-ination. In the case of low-force cold switching of RFMEMS switches in ambient air, typically a higher initialcontact resistance is observed [2], [25]. Contaminationaccumulated on the contact surface during cold switchingmight easily degrade the contact resistance by orders ofmagnitude.

2) Pure gold yields the lowest cycle number before electricalfailure. Solid-solution strengthened Au–Ni alloys showedan increased cycle number with an increased nickelcomposition. The tradeoff for the extended lifetimes isa relatively higher contact resistance (∼1.8 Ω). In theMEMS community, hot-switching failure is thought tobe primarily caused by arc-induced material transfer ortransient current heating. Mechanical wear is thereforethought to be a minor factor in the hot-switching failureprocess. However, our data indicate that strengtheningthe gold alloy substantially retards the contact failureprocess under hot-switch conditions. This might be anindicator that the arc/current-heating-induced materialtransfer could be affected by the mechanical propertiesof the contact materials. A more detailed investigation ofthe field-induced material transfer mechanisms is beingreported elsewhere.

3) Two-phase Au–Ni yields the largest cycle number whileproviding a stable and acceptable contact resistance(∼1 Ω). Compared with the solid-solution Au–Ni alloys,two-phase Au–Ni (20 at.% Ni) films have an interme-diate hardness (4.1 GPa) while yielding a much largercontact area due to the film smoothness. However, theintermediate hardness of this material does not fit ouroriginal expectation that the switching reliability shouldincrease with the hardness. Furthermore, the large contactarea is expected to enhance the material adhesion on thecontact surfaces, thus impacting the reliability. The ex-cellent performance of the two-phase alloy, as well as thecontrast to the trends displayed by the single-phase alloys,caused us to suspect surface-related effects. In order to in-vestigate the chemical differences between solid-solutionand two-phase sample surfaces, the chemical composi-tion of the sample surfaces was examined using XPS,comparing the Au–Ni (20 at.% Ni) solid-solution and theAu–Ni (20 at.% Ni) second-phase precipitation samples.

Fig. 14(a) shows the XPS spectrum of the solid-solutionAu–Ni (20 at.%) alloy. The nickel composition is calibratedbased on the Ni 2p peak. The nickel atomic percentage is∼11.7% (this value is only representative of the surface com-position). Fig. 14(b) shows the XPS spectrum for two-phaseAu–Ni (20 at.%). The calibrated nickel composition on thesurface is ∼30.8 at.% in this case. The high-resolution XPSspectrum on two-phase Au–Ni (20 at.%) is shown in Fig. 14(c).The dominant nickel species on the surface is identified as

Fig. 14. (a) XPS spectrum on Au–Ni (20 at.% Ni) solid-solution sample.(b) XPS spectrum on two-phase Au–Ni alloy sample (20 at.% Ni, after RTA).(c) High-resolution XPS spectrum on Au–Ni (20 at.% Ni, after RTA) two-phasesample. The listed binding energy is taken from [26]–[29].

Ni2O3 from the Ni 2p 3/2 position [28], [29]. Therefore, theprecipitation of nickel from solid-solution Au–Ni (20 at.%)alloys might be described as follows. Upon RTA, nickel startsto precipitate from the gold-rich matrix. Moreover, nickel nearthe surface precipitates preferentially at the surface (as well asat grain boundaries) and is oxidized upon air exposure. Theresults of the XRD and XPS analyses, taken together with theAu–Ni phase diagram, indicate that, after the RTA processing,some nickel atoms should remain as solid-solution hardeningadditions to the Au, some Ni exists as Ni-rich precipitates at thenanoscale, and some Ni migrates and forms nanoscale oxidizedNi2O3 islands on the surface. This engineered microstructure,having unique surface and volume compositions and properties,showed the best contacting performance among all the Au–Nisamples. These results point toward the need to understandand control both the following when selecting materials forMEMS contact switch applications: 1) material compositionand microstructure and 2) surface structure and properties.

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 8: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

294 JOURNAL OF MICROELECTROMECHANICAL SYSTEMS, VOL. 18, NO. 2, APRIL 2009

IV. CONCLUSION

In this paper, a newly developed switching degradation testfacility was used to evaluate Au–Ni alloys as contact materi-als for MEMS switches under hot-switching conditions. Theexperimental data shown in this paper provide a correlationbetween material properties (such as microhardness, electricalresistivity, surface topology, and surface microstructure) andmicrocontacting performance. In general, increasing the mate-rial hardness increases the switching reliability, compared withpure Au. However, surface roughness needs to be maintainedbelow a certain level in order to avoid detrimental effects oncontact resistance. It also has been shown that postdepositionRTA treatment results in a two-phase Au–Ni alloy, whichsignificantly improves the microcontact performance based onits unique material properties. The presence of some Ni2O3

oxide at the surface appears to correlate with a longer switchinglifetime. These results suggest that both material volume andmaterial surface properties must be considered for optimiz-ing the properties of alloy thin films for MEMS switchingapplications.

ACKNOWLEDGMENT

The authors would like to thank the many members of the Ex-treme Friction Multidisciplinary University Research Initiative(MURI) research group who contributed to this paper, includingA. Gruverman, D. Wu, M. Walker, and C. Brown.

REFERENCES

[1] C. Fung, “Industry study on issues of MEMS reliability and acceleratedlifetime testing,” in Proc. 43rd Annu. IEEE Int. Rel. Phys. Symp., 2005,pp. 312–316.

[2] B. D. Jensen, L.-W. Chow, K. Huang, K. Saitou, J. L. Volakis, andK. Kurabayashi, “Effect of nanoscale heating on electrical transport inRF MEMS switch contacts,” J. Microelectromech. Syst., vol. 14, no. 5,pp. 935–946, Oct. 2005.

[3] C. Brown, A. Morris, A. Kingon, and J. Krim, “Cryogenic performanceof RF MEMS switch contacts,” J. Microelectromech. Syst., vol. 17, no. 6,pp. 1460–1467, Dec. 2008.

[4] A. Granaldi and P. Decuzzi, “The dynamic response of resistive mi-croswitches: Switching time and bouncing,” J. Micromech. Microeng.,vol. 16, no. 7, pp. 1108–1115, Jul. 2006.

[5] P. Decuzzi, G. P. Demelio, G. Pascasio, and V. Zaza, “Bouncing dynamicsof resistive microswitches with an adhesive tip,” J. Appl. Phys., vol. 100,no. 2, p. 024 313, Jul. 2006.

[6] J. Oberhammer and G. Stemme, “Active opening force and passive contactforce electrostatic switches for soft metal contact materials,” J. Micro-electromech. Syst., vol. 15, no. 5, pp. 1235–1242, Oct. 2006.

[7] N. E. McGruer, G. G. Adams, L. Chen, Z. J. Guo, and Y. Du, “Mechanical,thermal, and material influences on ohmic-contact-type MEMS switchoperation,” in Proc. 19th MEMS, Istanbul, Turkey, 2006, pp. 230–233.

[8] R. A. Coutu, P. E. Kladitis, K. D. Leedy, and R. L. Crane, “Selectingmetal alloy electric contact materials for MEMS switches,” J. Micromech.Microeng., vol. 14, no. 8, pp. 1157–1164, Aug. 2004.

[9] R. A. Coutu, J. R. Reid, R. Cortez, R. E. Strawser, and P. E. Kladitis,“Microswitches with sputtered Au, AuPd, Au-on-AuPt, and AuPtCu alloyelectric contacts,” IEEE Trans. Compon. Packag. Technol., vol. 29, no. 2,pp. 341–349, Jun. 2006.

[10] D. Hyman and M. Mehregany, “Contact physics of gold microcontacts forMEMS switches,” IEEE Trans. Compon. Packag. Technol., vol. 22, no. 3,pp. 357–364, Sep. 1999.

[11] J. Schimkat, “Contact materials for microrelays,” in Proc. 11th MEMSWorkshop, 1998, pp. 190–194.

[12] D. J. Dickel and M. T. Dugger, “The effects of surface contam-ination on resistance degradation of hot-switched low-force MEMS

electrical contacts,” in Proc. 51st IEEE Holm Conf. Elect. Contacts, 2005,pp. 255–258.

[13] L. Chen, N. E. McGruer, and G. G. Adams, “Contact evolution in mi-cromechanical switch,” in Proc. WTC2005, World Tribology Congr. III,Paper WTC2005-63970. CD-ROM.

[14] X. Yan, N. E. McGruer, G. G. Adams, and S. Majumder, “Finite elementanalysis of the thermal characteristics of MEMS switches,” in Proc. 12thInt Conf. Transducers, Solid-State Sensors, Actuator Microsyst., 2003,vol. 1, pp. 412–415.

[15] Z. Yang, D. Lichtenwalner, A. Morris, S. Menzel, C. Nauenheim,A. Gruverman, J. Krim, and A. I. Kingon, “A new test facility for effi-cient evaluation of MEMS contact materials,” J. Micromech. Microeng.,vol. 17, no. 9, pp. 1788–1795, Sep. 2007.

[16] H. Okamoto and T. B. Massalski, Phase Diagrams of Binary Gold Alloys.Materials Park, OH: ASM Int., 1987, p. 196.

[17] A. S. Morris, S. Cunningham, D. Dereus, and G. Schropfer, “High per-formance integrated RF MEMS: Part 1—The process,” in Proc. 33rd Eur.Microw. Conf., 2003, pp. 21–24.

[18] M. Ohring, Material Sciences of Thin Films. New York: Academic,2002, p. 176.

[19] D. S. Perloff, “Four-point sheet resistance correction factors for thinrectangular samples,” Solid State Electron., vol. 20, no. 8, pp. 681–687,Aug. 1977.

[20] A. F. Crawley and D. J. Fabian, “A redetermination of the lattice spacingsand densities of gold–nickel solutions,” J. Inst. Met., vol. 94, pp. 39–40,1966.

[21] B. D. Jensen, L. W. Chow, R. F. Webbink, K. Saitou, J. L. Volakis, andK. Kurabayashi, “Force dependence of RF MEMS switch contact heat-ing,” in Proc. 17th IEEE Int. Conf. MEMS, 2004, pp. 137–140.

[22] S. C. Bromley and B. J. Nelson, “Performance of microcontacts testedwith a novel MEMS device,” in Proc. 47th IEEE Holm Conf. Elect.Contacts, 2001, pp. 122–127.

[23] R. S. Timsit, Electrical Contacts: Principles and Applications.New York: Marcel Dekker, 1999, ch. 1, pp. 3–6.

[24] O. Rezvanian, M. Zikry, C. Brown, and J. Krim, J. Microelectromech.Syst., submitted for publication.

[25] C. Brown, Physics Dept., North Carolina State Univ., Raleigh, NC,unpublished.

[26] V. I. Nefedov, M. N. Firsov, and I. S. Shaplygin, “Electronic structuresof MRhO2, MRh2O4, RhMO4 and Rh2MO6 on the basis of X-ray spec-troscopy and ESCA data,” J. Electron Spectrosc. Relat. Phenom., vol. 26,p. 65, 1982.

[27] K. S. Kim and N. Winograd, “X-ray photoelectron spectroscopic studiesof nickel–oxygen surfaces using oxygen and argon ion-bombardment,”Surf. Sci., vol. 43, no. 2, pp. 625–643, Jun. 1974.

[28] P. Lorenz, J. Finster, G. Wendit, J. V. Salyn, E. K. Zumadilov,and V. I. Nefedov, “ESCA investigations of some NiO/SiO2 andNiO–Al2O3/SiO2 catalysts,” J. Electron Spectrosc. Relat. Phenom.,vol. 16, p. 267, 1979.

[29] K. Kishi, “Adsorption of ethylenediamine on clean and oxygen cov-ered Fe/Ni(100) surfaces studied by XPS,” J. Electron Spectrosc.Relat. Phenom., vol. 46, p. 237, 1988.

Zhenyin Yang received the B.S. degree fromNanjing University, Nanjing, China, in 2001, and theM.S. degree in applied science from the College ofWilliam and Mary, Williamsburg, VA, in 2004. Heis currently working toward the Ph.D. degree in theDepartment of Materials Science and Engineering,North Carolina State University, Raleigh.

His research interests include thin-film processing,vacuum technology, microfabrication, and micro-engineering. His work continues with the develop-ment of RF microelectromechanical system switches

for mobile and satellite communications applications.

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.

Page 9: Comparison of Au and Au–Ni Alloys as Contact Materials for ...YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 289 when switch opening and

YANG et al.: COMPARISON OF Au AND Au–Ni ALLOYS AS CONTACT MATERIALS FOR MEMS SWITCHES 295

Daniel J. Lichtenwalner received the B.S. degreein metallurgical engineering from the Universityof Missouri, Rolla, in 1985, and the Ph.D. de-gree in materials science and engineering from theMassachusetts Institute of Technology, Cambridge,in 1990, under the guidance of Alfredo C. Andersonand David A. Rudman.

He was a Visiting Professor with the Depart-ment of Chemistry, Florida International University,Miami. He is with North Carolina State Univer-sity, Raleigh, where he was first a Postdoctoral Re-

searcher and is currently a Research Assistant Professor with the Department ofMaterials Science and Engineering. He has authored or coauthored more than50 journal articles, and holds one U.S. patent. His research interests includevacuum technology and thin-film processing, generally in the area of inorganicthin films for microelectronic applications.

Dr. Lichtenwalner is a member of the Materials Research Society and of theElectrochemistry Society. He is a Lifetime Member of Phi Kappa Phi.

Arthur S. Morris, III (M’87–SM’04) received theB.S. degree in physics and electrical engineering andthe M.S. and Ph.D. degrees in electrical engineer-ing from North Carolina State University (NCSU),Raleigh, in 1983, 1986, and 1993, respectively.

As a Scientist/Engineer concentrating on physicalelectronics and electromagnetic fields for more than20 years, he has contributed to device technolo-gies ranging from high-power traveling-wave tubesto millimeter-wave heterojunction bipolar transistorsand has developed products for markets from high-

voltage instrumentation to broadband communication systems. He was withMicrocosm Technologies (now Coventor) in 1999, where he led software andhardware development to support the transition of microsystems from thelaboratory into products for radio-frequency (RF) and optical applications. Heis currently Chief Technology Officer with wiSpry, Inc., Irvine, CA, developingthe next generation of highly integrated and agile RF products leveragingmicroelectromechanical systems and microstructures. He is also an AdjunctProfessor with NCSU. He has authored over 40 refereed papers and holdsseveral U.S. patents. His research interests include fundamental science andtechnology enabling high-volume commercial applications.

Dr. Morris is a member of Phi Kappa Phi, Eta Kappa Nu, andTau Beta Pi.

Jacqueline Krim received the B.A. (with honors)degree in physics from the University of Montana,Missoula, in 1978, and the Ph.D. degree in experi-mental condensed matter physics from the Universityof Washington, Seattle, in 1984, under the guidanceof J. Gregory Dash.

Before 1998, she was a Professor with the Depart-ment of Physics, Northeastern University, Boston,MA, after a year-long appointment as a NorthAtlantic Treaty Organization Postdoctoral Fellowwith the Université d’Aix–Marseille II, Marseille,

France. She is currently a Full Professor with the Department of Physics, NorthCarolina State University, Raleigh. She has authored numerous journal articlesand invited reviews for both scientific readerships and the general public.Her research interests include solid-film growth processes and topologies atsubmicrometer-length scales, nanotribology (the study of friction, wear, andlubrication at nanometer length and time scales), and liquid-film wettingphenomena.

Dr. Krim is a Fellow of the American Physical Society and the AmericanVacuum Society, and serves on several editorial and international advisoryboards in the area of tribology and nanoscale science.

Angus I. Kingon (M’91) was born in Pretoria,South Africa. He received the B.Sc. (Hons) degreein physical chemistry from the University of theWitwatersrand, Johannesburg, South Africa, in 1975,and the M.Sc. (cum laude) and Ph.D. degrees fromthe University of South Africa, Pretoria, the latterin 1981.

He held postdoctoral positions with PennsylvaniaState University, University Park, and with NorthCarolina State University (NCSU), Raleigh, during1981 and 1982. He spent a period of ten years with

the Council for Scientific Research, Pretoria, with the National Physical Re-search Laboratory, and, subsequently, with the National Institute for MaterialsResearch, where he was a Senior Chief Research Scientist and ProgramManager. In 1987, he was with the Department of Materials Science andEngineering, NCSU, as a Full Professor, where he held a joint appointment withthe Department of Management, Innovation, and Entrepreneurship, Collegeof Management. He is currently a Professor of engineering and the BarrettHazeltine University Professor of Entrepreneurship and Organizational Studieswith the Division of Engineering, Brown University, Providence, RI. He haspublished more than 300 papers in refereed journals, written six book chapters,and coedited seven books. He is an inventor and holds fifteen U.S. patents.

Dr. Kingon is a Fellow of the American Ceramic Society and is activein the Materials Research Society. He is a member of the FerroelectricsCommittee of the IEEE Ultrasonics, Ferroelectrics, and Frequency ControlSociety. He was the recipient of the Alcoa Research Award, the 2000 Ferro-electrics Achievement Award, and the Price Foundation Award as an InnovativeEntrepreneurship Educator for 2006.

Authorized licensed use limited to: North Carolina State University. Downloaded on May 13, 2009 at 14:49 from IEEE Xplore. Restrictions apply.