74
THE PENNSYLVANIA STATE UNIVERSITY SCHREYER HONORS COLLEGE DEPARTMENT OF ENGINEERING SCIENCE AND MECHANICS IMPLANTABLE MICROELECTRONIC DEVICES FOR MEASURING IN VIVO CORROSION OF MAGNESIUM ALLOYS BETH A BIMBER Fall 2011 A thesis submitted in partial fulfillment of the requirements for a baccalaureate degree in Engineering Science with honors in Engineering Science Reviewed and approved* by the following: Barbara A. Shaw Professor of Engineering Science and Mechanics Thesis Supervisor Charles E. Bakis Distinguished Professor of Engineering Science and Mechanics Honors Adviser Judith A. Todd P. B. Breneman Deparment Head Chair Professor, Department of Engineering Science and Mechanics Elizabeth Sikora Research Associate in Engineering Science and Mechanics * Signatures are on file in the Schreyer Honors College and Engineering Science and Mechanics Office.

BETH A BIMBER - ETDA

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: BETH A BIMBER - ETDA

THE PENNSYLVANIA STATE UNIVERSITY

SCHREYER HONORS COLLEGE

DEPARTMENT OF ENGINEERING SCIENCE AND MECHANICS

IMPLANTABLE MICROELECTRONIC DEVICES FOR MEASURING IN VIVO

CORROSION OF MAGNESIUM ALLOYS

BETH A BIMBER

Fall 2011

A thesis

submitted in partial fulfillment

of the requirements

for a baccalaureate degree

in Engineering Science

with honors in Engineering Science

Reviewed and approved* by the following:

Barbara A. Shaw

Professor of Engineering Science and Mechanics

Thesis Supervisor

Charles E. Bakis

Distinguished Professor of Engineering Science and Mechanics

Honors Adviser

Judith A. Todd

P. B. Breneman Deparment Head Chair

Professor, Department of Engineering Science and Mechanics

Elizabeth Sikora

Research Associate in Engineering Science and Mechanics

* Signatures are on file in the Schreyer Honors College and Engineering Science and

Mechanics Office.

Page 2: BETH A BIMBER - ETDA

i

ABSTRACT

Medical implants have been around for thousands of years. However, the idea of

biodegradable medical implants, for use as coronary stents or bone plates, is more recent. The

corrosion properties of magnesium, along with the fact that it is already needed in the human

body, make it a good candidate for a biodegradable implant material. However, pure magnesium

corrodes too fast, so this material needs to be alloyed.

Varying weight percent additions of titanium were alloyed with magnesium to create a

spectrum of physical vapor deposited Mg-Ti alloys. The alloys are electrochemically tested using

open circuit potential, polarization resistance, and EIS experiments to determine the specific

corrosion rates in vitro. The in vitro corrosion rates allow the varying alloys to be compared, to

determine the best corrosion rate. As much as the in vitro corrosion rate provides information

about how the alloy will act in a corrosive environment, in vivo corrosion data is needed. A

method for measuring the in vivo corrosion rates needs to be created.

An implantable microelectronic device with a working, counter, and reference electrode

on a micro-sized scale can solve the problem of how to measure the specific in vivo corrosion

rate. Currently, there are still issues to the specific processing in the manufacturing of said device,

such as the lift-off process of the working electrode. Further research will solve this problem,

since the implications of in vivo corrosion data can implemented in biodegradable coronary stents

and bone plates.

Page 3: BETH A BIMBER - ETDA

ii

TABLE OF CONTENTS

LIST OF FIGURES ................................................................................................................. iiv

LIST OF TABLES ................................................................................................................... vi

ACKNOWLEDGEMENTS ..................................................................................................... vii

Chapter 1 .................................................................................................................................. 1 Magnesium, Implants and Microelectronic Devices ........................................................ 1

Chapter 2 .................................................................................................................................. 3 Literature Review ............................................................................................................. 3 Magnesium ....................................................................................................................... 3

The Specific Reactions ............................................................................................. 3 Alloying Magnesium ........................................................................................................ 5

Titanium, Aluminum, and Yttrium as Alloying Elements ....................................... 6 In vitro versus In vivo ............................................................................................... 7

Previous Research within the Shaw Research Group ...................................................... 12

Thin Films and Vapor Deposition ............................................................................ 12 Determning the Corrosion Rates of Samles through Experimentation .................... 14 Previous Implantable Microelectronic Device Research ......................................... 16

Chapter 3 .................................................................................................................................. 18 Experimental Design and Methods .................................................................................. 18 In vitro Corrosion Experiments ........................................................................................ 18

Creating the Alloyed Magnesium Samples .............................................................. 19 Running PR and EIS Electrochemical Experiments ................................................ 23

Implantable Microelectronic Device ................................................................................ 25

Mask 1: Platinum-Titanium Layer (Reference and Counter Electrodes) ................. 26 Mask 2: Magnesium-Titanium Layer (Working Electrode) ..................................... 28 Mask 3: Parylene Coat ............................................................................................. 29

Chapter 4 .................................................................................................................................. 30 Results and Discussion ..................................................................................................... 30 Implantable Microelectronic Device ................................................................................ 30

Initial Three-Electrode Device ................................................................................. 30 Redesigned Three-Electrode Device ........................................................................ 31 Electrochemical Evaluation of Commercially Avaliable Titanium Rods ................ 35

Magnesium Aloy Samples for In vitro Testing ................................................................ 30 Sample Calculations for Corrosion Rate Determination .......................................... 48

Page 4: BETH A BIMBER - ETDA

iii Chapter 5 .................................................................................................................................. 51 Conclusions ...................................................................................................................... 51

Chapter 6 .................................................................................................................................. 52 Future Work ..................................................................................................................... 52

Appendix A .............................................................................................................................. 54

Appendix B .............................................................................................................................. 55

References ................................................................................................................................ 58

Page 5: BETH A BIMBER - ETDA

iv

LIST OF FIGURES

Figure 2-1: Set-Up for a basic battery or corrosion experiment ............................................. 4

Figure 2-2: Set-Up schematic for a sample that is both the anode and the cathode ................ 5

Figure 2-3: Pourbaix diagram for magnesium [8] .................................................................. 6

Figure 2-4: Histographic image of a magnesium stent in a vessel, 35 days after initial implantation [9] ................................................................................................................ 8

Figure 2-5: LAE442 magnesium sample (a), after 12 weeks in vivo (b), reconstructed from SRμCT images [14] ................................................................................................. 10

Figure 2-6: Pin with surrounding tissue (top), Optical Coherence Tomography (OCT) B-Scan images of magnesium sample and surrounding tissue [16] ..................................... 11

Figure 2-7: Physical Vapor Deposition (PVD) set-up schematic [23] .................................... 13

Figure 2-8: Nyquist plot for 1 mm/year corrosion rate [36] ................................................... 15

Figure 3-1: EB-PVD machine utilized to create the thin film magnesium alloy samples ....... 19

Figure 3-2: Magnesium alloy on silicon wafers, reflecting the Penn State logo [28] ............. 20

Figure 3-3: Wire connection to the future working electrode, for in vitro experiments ......... 21

Figure 3-4: Working electrode with an area masked for electrochemical testing, after five coats of PMMA [38] ........................................................................................................ 22

Figure 3-5: Schematic of the future working electrode immersed in an electrolyte ............... 22

Figure 3-6: Set-up for electrochemical experiments, with labels for the different electrodes ......................................................................................................................... 22

Figure 3-7: Another angle of the set-up for elctrochemical experiments, with labels for hot plate and temperature probe ....................................................................................... 22

Figure 3-8: LEdit drawings for the three initial masks, along with all three overlaid each other (scaled in mm) [39] ................................................................................................. 26

Figure 3-9: Wafer, after the photoresist has been exposed ..................................................... 27

Figure 3-10: A platinum-titanium layer that has been partially lifted-off............................... 28

Page 6: BETH A BIMBER - ETDA

v Figure 4-1: Initial three electrode with Mg-Ti working electrode that looks like swiss

cheese [images courtesy of Anthony Nacarrelli] ............................................................. 31

Figure 4-2: Front edge of the device, with the initial set of masks ......................................... 32

Figure 4-3: Front edge of the device, with the new, redesigned set of masks ......................... 32

Figure 4-4: LEdit drawings for the three redesigned masks (scaled in cm), along with all three masks overlaid ......................................................................................................... 33

Figure 4-5: Implantable microelectronic three electrode device with key features labeled .... 33

Figure 4-6: Uncleaved and uncoated device wafer, with titanium as the working electrode ........................................................................................................................... 34

Figure 4-7: Comparison between the old device and the new design ..................................... 35

Figure 4-8: Open circuit potential for a bulk titanium rod in HBSS at 37°C .......................... 36

Figure 4-9: SEM image of the redesigned microelectronic device, at 20x magnification ...... 38

Figure 4-10: ESEM images of corners and edges of a titanium working electrode device sample, with location of the image shown on the device by the red circle ...................... 39

Figure 4-11: Open circuit potential, polarization resistance, and EIS results for sample 144-3 (2.84 wt% Ti) in HBSS at 37°C ............................................................................. 42

Figure 4-12: Open circuit potential, polarization resistance, and EIS results for sample 144-4 (2.84 wt% Ti) in HBSS at 37°C ............................................................................. 43

Figure 4-13: Open circuit potential, polarization resistance, and EIS results for sample 145-4 (3.97 wt% Ti) in HBSS at 37°C ............................................................................. 44

Figure 4-14: Open circuit potential, polarization resistance, and EIS results for sample 145-6 (3.97 wt% Ti) in HBSS at 37°C ............................................................................. 45

Figure 4-15: Open circuit potential, polarization resistance, and EIS results for sample 145-7 (3.97 wt% Ti) in HBSS at 37°C ............................................................................. 46

Figure 4-16: Open circuit potential, polarization resistance, and EIS results for sample 146-4 (7.44 wt% Ti) in HBSS at 37°C ............................................................................. 47

Figure 4-17: ESEM image of the top of the magnesium sample 145-3 (3.97 wt% Ti) .......... 50

Figure 4-18: ESEM image of the top of the magnesium sample 145-3 (3.97 wt% Ti), at a lower magnification ......................................................................................................... 50

Figure 6-1: Image of a rat’s skull, with the implant location of the microelectronic implantable device (left), Skull protection cap on a rat ................................................... 52

Page 7: BETH A BIMBER - ETDA

vi

LIST OF TABLES

Table 3-1: List of samples, with weight % of titanium ........................................................... 18

Table 4-1: Open circuit potentials of titanium rods versus saturated calomel electrode (reference) in HBSS at 37 °C ........................................................................................... 37

Table 4-2: Open circuit potentials and corrosion rates (calculated from polarization resistance experiments) of titanium rods versus platinum foil (reference) in HBSS at 37°C ................................................................................................................................. 37

Table 4-3: Open circuit potentials and corrosion rates (via polarization resistance and EIS experiments) of various magnesium samples versus SCE (reference) in HBSS at 37°C ................................................................................................................................. 49

Page 8: BETH A BIMBER - ETDA

vii

ACKNOWLEDGEMENTS

I would like to thank these people for specifically helping me with my research and thesis

(no specific order):

i. Anna Hartsock – for all of her help with answering my many questions about the

experiments, assisting me with learning the in vitro corrosion testing procedure, and for

all of her help with everything else

ii. Daniel Cook – for helping me with taking SEM images of most of my samples

iii. Hitesh Basantani – for allowing me to follow him around to learn the recipe for making

the implantable microelectronic devices

iv. Derek Wolfe – for being an excellent lab partner and working on the connection issue

with the implantable microelectronic device

v. Anthony Naccarelli – for helping me perfect the process of creating in vitro corrosion

samples

vi. Robert Gresh – for physical vapor depositions for the magnesium alloys

vii. Dr. Elizabeth Sikora – for helping me to understand what my results meant and for

suggesting some key papers for my thesis

viii. Dr. Barbara Shaw – for allowing me the chance to work in her lab and for helping me out

when I was first interested in corrosion

ix. Dr. Mark Horn – for providing excellent knowledge on thin films and PVD materials

x. Dr. Bruce Gluckman – for providing his knowledge on implantable devices

xi. Scott Kralik – for teaching me how to run a Scanning Electron Microscope (SEM)

xii. Amber Black – for assisting me with imaging some of the PVD magnesium alloys

Page 9: BETH A BIMBER - ETDA

viii

xiii. Sean Pursel – for teaching me about the PVD system and how the magnesium thin films

are made

xiv. Materials Research Institute – for their knowledge, guidance, and use of their

nanofabrication facilities, which allowed me to actually create and fully understand what

I was working on.

Page 10: BETH A BIMBER - ETDA

1

Chapter 1

Magnesium, Implants, and Microelectronic Devices

Inventions are going towards a “greener” direction, becoming more environmentally

friendly. People are shying away from the plastics that seem to last forever and are instead using

chip bags and egg cartons that can degrade on their own. This same idea can be applied to

medical implants, specifically magnesium medical implants. They are implanted, but then

breakdown on their own, without any harm to the environment.

Magnesium is an alkaline earth metal that is the eighth most abundant metal in the earth’s

crust. It is not found in its pure state however, since it reacts with other elements (like oxygen) to

form a protective layer [1]. This protective layer actually aids in making the magnesium more

corrosion resistant. It is this property, the ability to improve the corrosion resistance by the

addition of other elements, which will be exploited for medical purposes.

The idea of metal medical implants is not new – the 17th century saw iron, bronze, and

gold being utilized for plates, sutures, and dental implants [2-3]. Today artificial heart valves,

knee replacements and hip joints, made from metals, ceramics, and polymers, are all being

implanted. The idea of biodegradable implant though, is relatively recent. A biodegradable

implant is one that will be implanted into the body, complete the necessary task, and then corrode

away. Magnesium is an excellent choice for this – it is the fourth most abundant mineral in the

human body [4]. However, this metal corrodes too quickly in order to complete the necessary task

of holding open a blood vessel for a few weeks.

The next step is to alloy the magnesium with a metal that is more corrosion resistant. This

new alloy needs to be tested in vitro, to determine the corrosion rate. However, the in vivo (inside

body) corrosion rates are orders of magnitude slower than the in vitro (outside of body) corrosion

rates. The aspects of in vivo corrosion testing need to be further researched; specifically, the

Page 11: BETH A BIMBER - ETDA

2

actual corrosion rates need to be obtained. An implantable device that is a tiny corrosion

experiment will be able to provide the exact corrosion rate of the sample. The device, on the

micro scale, needs to measure the corrosion rate, in real time, while still implanted, so that

corrosion data can be collected over the entire time of implantation. The main focus of this thesis

is the design and implementation of such a device so that in vivo corrosion rates can be collected

for this magnesium alloy to be utilized in future medical implants, like stents, bone plates, and

screws.

Page 12: BETH A BIMBER - ETDA

3

Chapter 2

Literature Review

As mentioned previously, metal implants have been utilized previously. Specifically,

magnesium as a corroding metal and as a biodegradable material (and stent) has been explored by

other researchers [5 – 7, 9 – 21]. The Shaw research group has also completed previous research

into the use of magnesium as a biodegradable stent [22 – 25].

Magnesium

Magnesium is an ideal matrix metal as an implant due to its specific properties. For

instance, magnesium has a high mechanical strength [5, 6] and fracture toughness [6]. Humans

have roughly 30 g of magnesium in their body. Most of this can be found in their bones. In fact,

the daily recommended dosage of magnesium is 420 mg for adult males and 320 mg for adult

females. Side effects with magnesium supplements are also rare [6]. The primary issue with the

use of magnesium in the body is that pure magnesium corrodes too quickly [5 – 7].

The Specific Reactions

The experiments to test the corrosion rate of the magnesium samples, specifically for the

use of medical implants, are run in Hank’s Balanced Salt Solution (HBSS), a simulated body

fluid (SBF). The recipe for creating the HBBS can be found in Appendix A. As stated by the

name, simulated body fluids attempt to replicate the natural salts and viscosity of human fluids, to

try to provide a similar environment for the experiments.

Page 13: BETH A BIMBER - ETDA

4

Once submerged in the SBF, electrochemical reactions occur on the metal surface. For

corrosion to begin, four items need to be present:

1) an anode

2) a cathode

3) a electrochemical connection between the anode and the cathode, and

4) an electrolyte

A battery set-up, like the one shown in Figure 2-1, is often imagined for this set-up.

Medical implants differ from this set-up in the fact that the anode and the cathode are

actually on the same surface, the surface of the implant. A simple schematic set-up is shown in

Figure 2-2.

connection

Anode Cathode

electrolyte

Figure 2-1: Set-Up for a basic battery or corrosion experiment

electrolyte

Page 14: BETH A BIMBER - ETDA

5

Magnesium is a reactive metal; when exposed to air, the oxygen forms a quasi protective

layer. When immersed in water, the overall reaction, along with the anodic and cathodic reactions

are shown below:

𝑀𝑔 + 2𝐻2𝑂 → 𝑀𝑔(𝑂𝐻)2 + 𝐻2

𝑀𝑔 → 𝑀𝑔2+ + 2𝑒−

2𝐻+ + 2𝑒− → 𝐻2

The corrosion rate of magnesium in distilled water is 0.6 mpy (mils per year, with 1 mil = 0.001

inch), which can be converted to 1.5 x 10-2 mm/yr [8].

Alloying Magnesium

Different elements can be alloyed with the magnesium in order to produce a more

desirable corrosion rate since magnesium has such a high corrosion rate. These alloys need to be

electrolyte

Figure 2-2: Set-Up schematic for a sample that is both the anode and the cathode

electrolyte

Anode

Cathode

connection

Equation 2-1

Equation 2-2

Equation 2-3

Page 15: BETH A BIMBER - ETDA

6

solid solution alloys, not magnesium with an alloy precipitate. The precipitate will only increase

the corrosion rate of the magnesium.

Titanium, Aluminum and Yttrium as Alloying Elements

The purpose of alloying an element with magnesium is to improve its surface oxide film

– like what occurred when magnesium was exposed to air and a quasi protective film was formed.

The Pourbaix diagram for magnesium is shown in Figure 2-3.

Pourbaix diagrams show over the range of pH’s and potentials that a metal will corrode,

be passive, or be immune to the surrounding environment [8]. Over the range of pH that is found

in the body (near a neutral pH of 7), magnesium is shown to corrode. When selecting an alloying

material the important factor centers around this pH range. A material that is passive in this

Figure 2-3: Pourbaix diagram for magnesium [8]

Page 16: BETH A BIMBER - ETDA

7

region would be most desired, like titanium. Current work in the Shaw Corrosion Group is based

on magnesium alloys containing varying levels of titanium, between 2 and 8 percent, by weight.

Titanium is a low density and good mechanical strength metal that is resistance to most

organic acids and chloride solutions. It is considered physiologically inert [26]. Titanium has

been utilized in other medical implant, like screws and hip joints.

Other metals that could improve the corrosion rate in magnesium, besides titanium, are

aluminum and yttrium. Aluminum is a silver-white metal that is light, generally non-toxic, non-

magnetic, and non-sparking. It is malleable and easily cast [27]. Yttrium is found in lathanoid

rare-earth minerals. It is silvery-metallic in color [28].

The issue of cytotoxicity also needs to be researched before any implantation to test the

in vivo corrosion rates can be completed. The rare earth elements, like yttrium, do not have much

cytotoxicity research. Frank Feyerabend and et al. [29] completed experiments into the

cytotoxicity of magnesium alloyed with rare earth elements.

In vitro versus In vivo

Witte [7] mentioned that similar alloying magnesium samples produced different

corrosion rates based on an in vitro (outside of body) or an in vivo (inside body) experimental

setting. The difference is mentioned as being close to four orders of magnitude. This comparison

was made between in vitro corrosion rates and synchrotron-radiation based microtomography

images to determine the volume loss of the implanted specimen.

However, the corrosion rates for in vivo experiments rarely are put into specific numbers.

This ultimately means that in vivo corrosion experiments will need to be conducted, to insure that

the corrosion rates obtained in vivo are similar or on the same scale as the in vitro experiments

conducted for the past few years within this corrosion group.

Page 17: BETH A BIMBER - ETDA

8

Other research groups; however, have conducted in vivo corrosion experiments. Heublein

et al [9], produced magnesium alloy (AE21, which contains 2% aluminum and 1% rare earth)

stents, which were later implanted into domestic (German) pigs. The goal of this experiment was

to see, through histographic images, like the one shown in Figure 2-4, the degradation of the

stent. The white in the figure is the magnesium stent. The pink is the domestic pig coronary

artery. The target was to have 50% weight lose by the sixth month mark. The conclusion that was

drawn stated that magnesium alloyed stents were “a realistic alternative to permanent implant:

[9].

Xu and et al [10] examined magnesium alloys coated with calcium phosphate through a

phosphating process by an in vivo corrosion test. A magnesium rod sample was implanted into the

femur of a rat. Histographic analysis and residual area calculations provided percent degradation

of the sample. The byproducts of the corrosion reactions were carefully examined and recorded

throughout the experiment, and it was discovered that the calcium phosphate coating actually may

Figure 2-4: Histographic image of a magnesium stent in a vessel, 35 days after initial implantation [9]

Page 18: BETH A BIMBER - ETDA

9

have promoted early bone growth at the interface between the bone and magnesium alloy [10]. It

was also noted that magnesium was an excellent candidate for bone implants because the material

properties of magnesium were similar to bone [10 - 11].

Zhang and et al researched the degradation of the magnesium alloy, the response of the

bone to the magnesium alloy, and the blood composition (since the magnesium will degrade)

[12]. Using light microscopy and scanning electron microscopy, the samples showed differences

in the rates of degradation between implants in the marrow channel and implants in the cortical

bone. The samples that were in the marrow channel degraded faster. It was also observed that

bone started growing around the magnesium alloy at around six weeks, and though there was a

change in the blood composition, there was no disorder in the liver or kidneys [12].

Witte and group [13-21] researched into the use of magnesium in bone implants. In one

experiment, four different magnesium alloys (AZ31, AZ91, WE43, and LAE442) and a control

polymer rod were implanted into the femura of guinea pigs. After six and eighteen weeks,

histomorphologic images were generated. The corroded samples were analyzed through SEM

images, element mapping and X-ray diffraction to determine the degradation of the samples. The

conclusion of their experiment was that high magnesium concentration could lead to bone cell

activation, since the bone closest to the magnesium alloys had a higher bone mass [13]. In another

experiment, also by Witte [14], the specific magnesium alloy LAE442 (4% weight lithium, 4%

weight aluminum, and 2% weight rare earth metals) was tested by being implanted into the

medial femur condyle of adult rabbits. Non-coated LAE442 rods and rods coated in MgF2 were

investigated. The deductions from these trials were that the LAE442 magnesium alloys had low

corrosion rates, reacted favorably in a bioenvironment, and that samples coated in MgF2 had even

lower corrosion rates than uncoated samples [14]. Figure 2-5 shows the images of the actual

samples and then a reconstructed image (through synchrotron-radiation-based micro-computed

tomography (SRμCT)), which illustrates the pitting corrosion seen on this sample.

Page 19: BETH A BIMBER - ETDA

10

One problem that Witte [15] had to overcome was the fact the in vivo corrosion rates

were challenging to calculate. Pictures were taken throughout the experiment in order to see the

magnesium alloy corrode [15]. Figure 2-6 shows a titanium pin, by a histographic image, and an

in vivo image taken through optical coherence technology [16].

Figure 2-5: LAE442 magnesium sample (a), after 12 weeks in vivo (b), reconstructed from SRμCT images [14]

Page 20: BETH A BIMBER - ETDA

11

Witte also completed extensive research into the use of magnesium of a medical

biomaterial [17 – 18]. The biocompatibility of magnesium was seen when magnesium wires were

used as ligatures to stop bleeding vessels in 1878 [17]. Since then, various experiments have

determined that it is the salts in blood increase the corrosion rate of the magnesium compared to

the corrosion of magnesium by water, and that thickness and location of the magnesium

determine the corrosion rate. Magnesium samples placed in areas with high blood flow had

higher corrosion rates and it was noted that corrosion rates were different across species and

tissues. It was also seen that magnesium corrosion products were hydrogen and nitrogen gas (in

vivo) that appeared inert and were gradually absorbed [17 – 18].

Figure 2-6: Pin with surrounding tissue (top), Optical Coherence Tomography (OCT) B-Scan images of magnesium sample and surrounding tissue [16]

Page 21: BETH A BIMBER - ETDA

12

The experiments discussed previously, by groups such as Heublein [9], Xu [10], Zhang

[12], and Witte and group [13-21] prove that magnesium is a feasible biodegradable material. The

issue is that magnesium’s high corrosion rate needs to be lowered; one way to accomplish the

lower corrosion rate is to alloy the magnesium. However, future research into the specific in vivo

corrosion rate would provide further insight.

Previous Research within the Shaw Research Group

Thin Films and Vapor Deposition

Previous research has shown that the corrosion rates of bulk magnesium, even bulk

magnesium alloys, are too high. One way to slow down the corrosion rate is by vapor depositing

the specific materials onto a wafer. The process of physical vapor deposition allows the alloying

element to be scattered through the metal (magnesium) matrix. This extends the passivity-

enhancing element further into the matrix, thereby enhancing the corrosion resistance of the alloy.

The corrosion rate benefits of using physical vapor deposition (PVD) magnesium as opposed to

bulk magnesium were extensively research by Wolfe [22], Petrilli [23], and Van Pelt [24].

Wolfe noticed that the “microstructure and surface morphology of the magnesium alloy

on the silicon wafer was modifiable during physical vapor deposition process through the

variation of evaporation power, pressure, temperature, ion bombardment, and the source-to-

substrate distance” [22]. Petrilli also looked into the better corrosion rates, specifically of the

magnesium alloy WE43, and the in vitro cytotoxicity [23]. Van Pelt [24] looked more specifically

at the magnesium-aluminum alloys (AZ series). Magnesium alloyed with aluminum had a lower

corrosion rate than pure magnesium, though it is noted that the corrosion rate was also improved

Page 22: BETH A BIMBER - ETDA

13

from a bulk sample by the PVD process. The method of PVD also allows for higher

concentrations of aluminum to be alloyed with the magnesium [24].

For the thin films utilized by this group, the process of physical vapor deposition (PVD)

is done through electron beam evaporation [30]. Electron guns are aimed at the feed-stock

materials in a vacuum chamber. The materials (specifically magnesium and titanium) are heated

until they evaporate. The materials then condense on the silicon single crystal wafer positioned in

the top of the chamber. The wafer is continuously rotating, so that an even surface of condensed

magnesium and titanium can grow upon the said wafer. Figure 2-7 shows a schematic of the dual

gun electron beam physical vapor deposition (EB PVD) machine used to deposit the alloys in this

thesis. Several of the important elements in this figure are labeled.

Figure 2-7: Physical Vapor Deposition (PVD) set-up schematic [23]

Crucibles for holding materials

Roughing pump

Rotating substrate holder

Cryopump

Page 23: BETH A BIMBER - ETDA

14

Determining the Corrosion Rates of Samples through Experiments

The corrosion rates of the magnesium and titanium alloys deposited upon the silicon

wafers can be found through the electrochemical techniques of polarization resistance and

electrochemical impedance spectroscopy. During polarization resistance experiments, the

resistance of the sample to corrode is measured, as the potential of the sample is forced through

the open circuit potential (from anodic to cathodic).

During polarization resistance experiments, the potential of the magnesium is forced

away from the corrosion potential (OPCP). The OCP is the equilibrium voltage of the material

when it will corrode. This then causes current to flow between the electrodes and drives the

electrochemical equations. The polarization resistance (Rp), or the resistance of the metal to

corrosion, can then be found using the following equation:

𝑅𝑝 = �∆𝐸∆𝑖�∆𝐸→0

where ΔE is the change in the applied potential around the corrosion potential and Δi is the

resulting polarization current density.

The current density during the reaction is based upon the kinetics of the corrosion

reaction and the diffusion of the reactants towards and away from the electrode. This

phenomenon can be written as the Bulter-Volmer equation:

𝑖 = 𝑖𝐶𝑂𝑅𝑅 �𝑒�2.303 𝜂

𝑏𝑎� − 𝑒�−2.303 𝜂

𝑏𝑐��

where i is the measured current density, iCORR is the corrosion current density , ba and bc are the

Tafel slopes, and η is the overpotential, which can be written as:

𝜂 = 𝐸 − 𝐸𝐶𝑂𝑅𝑅

where E is the applied potential and ECORR is the corrosion potential. The corrosion current

density and the Tafel slopes are obtained from the experimental data [31 - 32].

Equation 2-4

Equation 2-5

Equation 2-6

Page 24: BETH A BIMBER - ETDA

15

For overpotentials close to the corrosion potential, Equation 2-5 can be rewritten as:

𝑖𝐶𝑂𝑅𝑅 = 2.303 � 𝑏𝑎𝑏𝑐𝑏𝑎+𝑏𝑐

� � 1𝑅𝑝�

Using Faraday’s Law, the corrosion rate can then be calculated:

𝐶𝑜𝑟𝑟𝑜𝑠𝑖𝑜𝑛 𝑅𝑎𝑡𝑒 = 𝐾 𝑖𝐶𝑂𝑅𝑅𝜌

𝐸𝑊

where ρ is the alloy density [g/cm3], EW is the equivalent weight [g/equivalent], iCORR is the

corrosion current density [µA/cm2], and K is the constant for converting units [32 - 35].

In electrochemical impedance spectroscopy (EIS) experiments, a small amplitude AC

current is applied to the sample. This method can provide information about the corrosion rate,

along with electrochemical mechanisms, reaction kinetics, and can detect localized corrosion

[36]. Localized corrosion is where the corrosion occurs at a specific (localized) small area.

The polarization resistance, Rp, can be found using the Nyquist plot produced from this

experiment. An example of this plot can be found in Figure 2-8.

Figure 2-8: Nyquist plot for 1 mm/year corrosion rate [36]

Equation 2-7

Equation 2-8

RΩ RΩ+ Rp

Page 25: BETH A BIMBER - ETDA

16

From the graph, RΩ is the resistance of the solution, and Rp is the polarization resistance. The

polarization resistance that is found experimentally can then be utilized in Equations 2-7 and 2-8

in order to determine the corrosion rate of the sample.

Previous Implantable Microelectronic Device Research

The issue between the in vitro and the in vivo corrosion rates needs to be addressed, to

ensure that at least the experiments performed in vitro have some bearing on the actual corrosion

rates in vivo. The fact that some research groups have found differences of four orders of

magnitude [7] causes some concern.

Specific in vivo corrosion rates are not listed. Therefore, the only way to actually

determine the in vivo corrosion rates is to perform corrosion experiments while a sample is

physically in a host. This is to be completed by implanting miniature three-electrode devices into

a rat, and running the polarization resistance and EIS experiments on the samples. Pytel [25]

developed a set of three masks and a manufacturing process that created this microelectronic

implantable device.

The purpose of the microelectronic implantable device is to run corrosion experiments

and to be able to collect the same data as from the macro sized experiments. This means that the

same set-up needs to be present. The reference electrode, which the corrosion of the alloy is

compared back to, is platinum. The counter electrode, also known as the auxiliary electrode, is

also platinum. The counter electrode is the cathode to the samples’ anode.

Platinum is an inert material in terms of corrosion [37]. A simple reference electrode was

needed for the microelectronic implantable device, so it is being explored as a possible reference

electrode. The fact that platinum is very inert makes it an excellent reference electrode. Since the

corrosion rates will be based on the platinum electrode, the circuit potential of the platinum must

Page 26: BETH A BIMBER - ETDA

17

remain steady. In fact, this circuit potential must remain steady for long periods of time, since if it

is utilized as the counter reference on the device, it needs to last for at least a few weeks. De

Haro, et al. discovered that platinum with good adhesion was crucial to low corrosion rates [37] –

a property that is necessary to maintain a steady circuit potential.

Pytel [25] worked on the mask set and the manufacturing process of the microelectronic

implantable device. The microelectronic implantable device needs to be finalized so that in vivo

corrosion rates can be measured. The device will be implanted into the skull of a rat, to measure

the in vivo corrosion rate. This will provide needed corrosion rates of specific magnesium alloys

that can be utilized for biodegradable stents, bone plates, and other medical implants.

Page 27: BETH A BIMBER - ETDA

18

Chapter 3

Experimental Design and Methods

In vitro Corrosion Experiments

As mentioned previously, alloyed magnesium samples show promise in being

biodegradable materials for medical implants. The weight percentage of specific alloying

element(s) to create the precise magnesium alloy with the desired corrosion rate still needs to be

found. This can be done by experimenting, in vitro, with various magnesium alloys; the in vitro

data will provide general guideline for the in vivo corrosion rate.

The samples being tested are crafted from 99.98% pure magnesium and commercially

pure titanium, with varying alloying additions of titanium. Table 3-1 shows a listing of samples

utilized in the electrochemical experiments and their weight percentage of titanium in the alloy

that was deposited on a wafer.

The corrosion rates are based upon the polarization resistance and electrochemical

impedance spectroscopy (EIS) experiments. The physical set-up for both polarization resistance

and EIS are the same. However, before the experiment can be conducted, an electrochemical test

sample needs to be made.

Table 3-1: List of samples, with weight % of titanium Sample (Wafer) Number % wt. Titanium

144 2.84 145 3.97 146 7.44

Page 28: BETH A BIMBER - ETDA

19

Creating the Alloyed Magnesium Samples for In vitro testing

The samples for the in vitro testing are made using a dual gun electron beam physical

vapor deposition process. Figure 3-1 shows the specific machine that is utilized to create the

samples.

The thin film is created by evaporating the metals, held by crucibles at the bottom

chamber. The evaporated metals condense on the surface of the wafer, which is rotating at the top

of the chamber. The EB-PVD process creates many layers of tiny grains of the solid solution

Figure 3-1: EB-PVD machine utilized to create the thin film magnesium alloy samples

Page 29: BETH A BIMBER - ETDA

20

metal alloy. In the case of Mg-Ti alloy, it improves the corrosion rate. The magnesium alloys are

deposited on silicon wafers, as shown in Figure 3-2.

The magnesium-titanium alloy deposited wafers are stored in a desiccator until use. The

entire wafer cannot be used for just a single electrochemical experiment since only a couple of

wafers can be processed in a day. Therefore, the wafer is cleaved into pieces; experiments can be

tested from the same deposit. Some pieces are also utilized for ICP (inductively coupled plasma),

so that the composition of the magnesium sample can be found. Other samples are utilized for

SEM imaging, so that the morphology of the deposit can be determined.

The remaining part of the wafer was cleaved into samples and processed for

electrochemical testing. An electrical connection (a wire) attached, so that there is an electrical

connection between the working electrode (the sample that will be corroded) and the potentiostat,

so that electrochemical data can be collected during the experiment. A common coated electrical

wire was attached to the alloy using a conductive sliver epoxy (EPO-TEK® H22). This step

Figure 3-2: Magnesium alloy on silicon wafers, reflecting the Penn State logo [28]

Page 30: BETH A BIMBER - ETDA

21

provided an electrical connection to the magnesium alloy and allowed for the sample to be safely

connected to the potentiostat. A safe and reliable connection to the potentiostat, via alligator clips

and cables, is needed to collect electrochemical data during the experiment. Figure 3-3 shows the

attached wires bonded to the PVD magnesium alloy on the silicon wafer.

Since a known area is needed for proper electrochemical testing, an area of the sample

was left unmasked. The silicon wafer, the wire, and the edges of the wafer and the magnesium-

titanium thin film also need to be protected. Smaller test areas are better, since there is a lower

likelihood that major defects in the thin film are present.

The electrochemical sample was masked with PMMA. PMMA, poly(methyl

methacrylate), is a transparent thermoplastic. The masking ensures that the silicon wafer and the

electrical connection would not interfere with the experiment. The alloy that was to be tested was

not coated. The PMMA is also coated up over the electrical connection, onto the wire, and over

the coated section of the wire (approximately 3-4 cm from the epoxy connection and at least 1 cm

up the insulated wire coating), to ensure there is no interaction from the metal wire. Five coats of

Figure 3-3: Coated wire connection to the working electrode, for in vitro experiments

Page 31: BETH A BIMBER - ETDA

22

PMMA ensure that all of the areas that cannot be affected by corrosion are completely covered.

Figure 3-4 shows what the future working electrode looks like after five coats of PMMA.

Figure 3-5 shows a schematic of how the sample will be immersed in the simulated body

fluid electrolyte. The silicon wafer and the edges of the PVD magnesium alloy are now protected

from corrosion, and a finite test area is open to the electrolyte.

Figure 3-4: Working electrode with an area masked for electrochemical testing, after five coats of PMMA [38]

Figure 3-5: Schematic of the future working electrode immersed in an electrolyte

Page 32: BETH A BIMBER - ETDA

23

Running PR and EIS Electrochemical Experiments

The set-up for both the polarization resistance and electrochemical impedance

spectroscopy experiments is the same. The samples are run in simulated body fluids (Hank's

Balanced Salt Solution) at body temperature (37°C) (recipe for HBSS can be found in Appendix

A). All of the experiments are conducted in this electrolyte at this temperature, to mimic the

actual biological environment.

The three electrodes are the working, the counter, and the reference electrode. The Mg-Ti

sample is the working electrode; platinum foil is the counter electrode; the saturated calomel

electrode is the reference electrode. A thermometer measures the temperature of the electrolyte.

The metal thermometer is kept in a glass vial so that the metal of the thermometer does not

interact with the corrosive electrolyte. Figure 3-6 and Figure 3-7 show this set-up, with the

connections to the Gamry potentiostat and its computer controller.

Page 33: BETH A BIMBER - ETDA

24

Figure 3-6: Set-up for electrochemical experiments, with labels for the different electrodes

Figure 3-7: Another angle of the set-up for electrochemical experiments, with labels for hot plate and temperature probe

SCE (reference electrode)

Magnesium sample

(working electrode)

Platinum (counter electrode)

Temperature Probe

Hot plate, for temperature control

Page 34: BETH A BIMBER - ETDA

25

The data collected from the electrochemical experiments is then used in equations 2-4 through 2-

8 to determine the corrosion rate of the magnesium alloy. The polarization resistance in these

equations can be obtained from either a linear polarization experiment or an EIS experiment.

Implantable Microelectronic Device

Running corrosion experiments in vivo requires three electrodes and will be on the micro

scale. Current nanotechnology manufacturing processes allow for precise multilayered structures

to be manufactured, which is necessary for creating a three electrode, implantable microelectronic

device for measuring the corrosion rate of implantable metals. A more detailed recipe for the

manufacturing of the implantable microelectronic device is listed in Appendix B.

There are three masks needed to create the implantable device, using photolithography

processing. Figure 3-8 shows the three individual masks, along with all three masks juxtaposed.

Page 35: BETH A BIMBER - ETDA

26

Mask 1: Platinum-Titanium Layer (Reference and Counter Electrodes)

Mask 1 allows for a pattern of platinum and titanium to be deposited on the wafer. This is

done by spinning a photoresist assist (LOR2A) and a photoresist (S1813) onto the six-inch

oxidized silicon wafers. The photoresist assist allows for complete lift-off of the photoresist, and

subsequently the platinum and titanium layers on top, which is necessary to insure that no

platinum or titanium remains except in the desired areas.

Figure 3-8: LEdit drawings for the three initial masks, along with all three overlaid each other(scaled in mm) [39]

MASK 1 MASK 2

ALL 3 MASKS MASK 3

Page 36: BETH A BIMBER - ETDA

27

A Karl Suss MA/BA6 was utilized to align the mask and the wafer, and to also expose

the photoresist-covered wafer to UV light. The UV light breaks the chemical bonds between the

polymer chains in this positive photoresist, making the areas exposed to the light soluble. The

exposure tool did not have the correct mask holder size necessary for the complete exposure of an

entire wafer, so the wafers were divided into quarters. Processing of the devices will be

conducted by each quarter wafer from this point onward. The wafers were then soaked in a

developer tool (CD-26) to remove the exposed photoresist. Figure 3-9 shows a wafer after the

photoresist has been exposed. The darker portion of the wafer is the exposed photoresist.

The Kurt J. Lesker CMS System at the Materials Research Laboratory (MRL) sputtered

thin layer of titanium and platinum onto the wafers and the patterned photoresist. The wafer is

Figure 3-9: Wafer, after the photoresist has been exposed

Page 37: BETH A BIMBER - ETDA

28

placed in an ultrasonic acetone bath and cleaned until the platinum-titanium layers were removed

from the substrate. Figure 3-10 shows the platinum-titanium layer during the lift-off process.

Mask 2: Magnesium-Titanium Layer (Working Electrode)

The purpose of mask 2 is to create the working electrode. The PVD process mentioned in

“Thin Films and Vapor Deposition” section in chapter 2 is the method used to deposit the Mg-Ti

alloy upon the wafer.

After the Mg-Ti alloy has been vapor deposited on the wafer, photoresist (S1813) is spun

on top. Mask 2 is aligned, and the photoresist is exposed to UV light. Once again, the exposed

resist is placed in a developer bath (CD-26) to remove the resist that was altered by the light.

Figure 3-10: A platinum-titanium layer that has been partially lifted-off

Page 38: BETH A BIMBER - ETDA

29

The wafer was then placed in a HNO3:H2O bath to remove the unwanted magnesium-

titanium. Acetone and IPA were used to rinse the wafer, to insure that the wafer was cleaned and

the unwanted magnesium-titanium was removed.

Mask 3: Parylene Coat

Mask 3 insures that all of the electrodes are exposed, and that the rest of the device is

coated in parylene-C (only one chloride group, is approved for biomedical applications) for

protection from the HBSS (for in vitro testing) and blood (in vivo testing). Parylene is vapor

deposited onto the wafer, in order to achieve at least a coating thickness of 0.5 μm for protection.

Photoresist (S1813) is spun onto the wafer, exposed, and then developed (CD-26). An oxygen

plasma etcher (PT720 machine) was then utilized to remove the parylene. The areas that were not

exposed to light were protected from the plasma by both parylene and photoresist. The plasma

was 90% O2 and 10% Ar. The wafers were etched in 30 second intervals, until the parylene was

removed.

The wafer still needs to be cleaved into individual devices, and the edges need to also be

coated with PMMA. The final product is an implantable device, which has a parylene coating,

with exposed electrodes and partially exposed leads.

Page 39: BETH A BIMBER - ETDA

30

Chapter 4

Results and Discussion

Implantable Microelectronic Device

Initial Three-Electrode Device

A silicon wafer of microelectronic devices, with magnesium-titanium working electrodes,

was created, in order to start in vitro testing. However, after the wet etching of the working

electrode (mask 2), it was seen that the magnesium-titanium alloy was already significantly

corroded.

Figure 4-1 is an optical microscope image of the device wafer revealing that overetching

of the Mg-Ti working electrode exposed a thin layer of underlying platinum at the edge of the

Mg-Ti layer. The exposure of platinum creates a galvanic cell between the Mg-Ti alloys and the

Pt. This in essence results an unwanted battery where the Pt is the cathode and the Mg-Ti

becomes the anode. The wet etch step in with the original masks is a fatal flaw in the processing

and necessitates the redesign of the masks.

Page 40: BETH A BIMBER - ETDA

31

Redesigned Three-Electrode Device

In addition to correcting the fatal flaw, the masks were improved in several other ways to

make the device smaller and allow for easier electrical connection. Adding tolerances around the

outside of mask 2 and mask 3 were used to address the over etching issue. This will mean that the

edges of the working electrode will now cover the platinum underneath the working electrode and

the protective parylene coat will make sure to cover the edges of the working electrode. Figure 4-

2 shows the front edge of the device (the cross section through the device), using the initial set of

masks. It can clearly be seen in the eighth image in this figure, how the platinum layer was

exposed under the magnesium-titanium working electrode. Figure 4-3 again shows the front edge

of the device (the cross section through the device), with the redesigned masks. In the eighth

image the working electrode layer now extends over the platinum underlayer. The mask sets were

designed using LEdit and were created at the Pennsylvania State University.

Figure 4-1: Initial three electrode device with Mg-Ti working electrode that looks like swiss cheese [images courtesy of Anthony Nacarrelli]

Mg-Ti working electrode

Pt

Page 41: BETH A BIMBER - ETDA

32

Figure 4-2: Front edge of the device, with the initial set of masks

Figure 4-3: Front edge of the device, with the new, redesigned set of masks

1 2 3 4 5 6 7 8 9 10 11 12

1 2 3 4 5 6 7 8 9 10 11 12

Page 42: BETH A BIMBER - ETDA

33

Figure 4-4 shows the redesigned mask layer set, also created using LEdit. Figure 4-5 shows the

new set of mask (each individual masks, along with all three juxtaposed), with a dimension

legend to show the relative size of the microelectronic device. Figure 4-5 labels the different

electrodes of the redesigned device.

Figure 4-4: LEdit drawings for the three redesigned masks (scaled in cm), along with all three masks overlaid

Figure 4-5: Implantable microelectronic three electrode device with key features labeled

Mg-Ti working electrode Pt reference electrode Pt counter electrode

Electrical connection pad

MASK 1 MASK 2

ALL 3 MASKS MASK 3

Page 43: BETH A BIMBER - ETDA

34

After the redesigned set of masks was completed, it was discovered that the filament on the second gun in the physical vapor deposition chamber was broken, which meant that only one metal could be deposited at a time. This provided the opportunity to test the design of the three electrode device with a more forgiving working electrode, a material with a low corrosion rate that is commonly utilized in the body. Titanium replaced a magnesium-titanium alloy as the working electrode. Since titanium was used, a hydrofluoric acid, rather than nitric acid, was employed to

remove the excess working electrode. The third mask, which is the parylene coat, was left off due

to time constraints – instead, for testing, the individual devices were cleaved and then coated with

PMMA to cover the edges, back, and between the electrodes of the wafer. Figure 4-6 shows the

uncleaved and uncoated new device wafer, with titanium as the working electrode.

Figure 4-7 shows the old device (though unfinished) compared to the new design. As

mentioned previously, there is a difference in size, the new device is about half the size of the

initial three electrode device. In addition there is a tolerance around the working electrode and

Figure 4-6: Uncleaved and uncoated device wafer, with titanium as the working electrode

Page 44: BETH A BIMBER - ETDA

35

parylene so that the reference electrode is not exposed during the wet etch process and the

electrical contacts have more space between them.

The next stage in the research is to test the new second generation device. Since the

device was completely redesigned, testing needs to show that the corrosion rates measured for the

working electrode on this new smaller device are comparable to results obtained from a macro-

sized working electrodes.

Electrochemical Evaluation of Commercially Available Titanium Rods

The first set of titanium trials were conducted on commercially pure titanium rods. The

purpose of the titanium rod experiments was to check the open circuit potential against varying

reference electrodes. The rods were sanded by 200 and 600 grit sandpaper before each trial. In

these experiments, the working electrode was the titanium and the reference electrode changed

between each set of trials. This was to insure that the titanium samples exhibited repeatable

Figure 4-7: Comparison between the old device and the new design

Page 45: BETH A BIMBER - ETDA

36

results (with the saturated calomel electrode) and the results that could be compared with the

device (when platinum was utilized as the reference electrode).

The first set of experiments consisted of measuring open circuit potential versus time

experiments were conducted on the titanium rods, with the saturated calomel electrode as the

reference electrode. The open circuit potential measures the thermodynamic tendency of the

sample to corrode. An example of the plot that was generated is shown in Figure 4-8, which is

specifically for titanium rod sample #1, trial #1 (versus a saturated calomel electrode).

Figure 4-8: Open circuit potential for a bulk titanium rod in HBSS at 37°C

-0.55

-0.53

-0.51

-0.49

-0.47

-0.45

-0.430 300 600 900

OCP

vs S

CE (V

)

Time (seconds)

Page 46: BETH A BIMBER - ETDA

37

Table 4-1 shows the results of the measured open circuit potential (OCP) of commercially pure

titanium rods in HBSS at body temperature. These experiments also provide a basis for future

corrosion testing of the bulk titanium rods against another reference electrode (like platinum foil).

The device is based on a platinum reference electrode; in order to better compare results

across varying reference electrodes, the experiments need to be run against different reference

electrodes. The experiments with the platinum reference electrode showed more variability than

the experiments conducted with the saturated calomel electrode. The experimental results are

shown in Table 4-2.

Table 4-1: Open circuit potentials of titanium rods versus saturated calomel electrode (reference) in HBSS at 37°C

Ti versus SCE OCP Trial 1 (V) OCP Trial 2 (V) OCP Trial 3 (V)

Ti Bulk sample 1 -0.4391 -0.4932 -0.4624 Ti Bulk sample 2 -0.4285 -0.5129 -0.5328 Ti Bulk sample 3 -0.4719 -0.5108 -0.4791

Table 4-2: Open circuit potentials and corrosion rates (calculated from polarization resistance experiments) of titanium rods versus platinum foil (reference) in HBSS at 37°C

Ti versus Pt foil

CR Trial 1 CR Trial 2

OCP Trial

1 OCP Trial

2 mpy mm/day mpy mm/day Ti Bulk sample 1 -0.5925 -0.7012 1.31E-01 9.12E-6 1.48E-01 10.3E-6 Ti Bulk sample 2 -0.3326 -0.1958 1.19E-01 8.28E-6 1.40E-01 9.74E-6 Ti Bulk sample 3 -0.6940 -0.5560 3.52E-02 2.45E-6 8.60E-02 5.98E-6

Page 47: BETH A BIMBER - ETDA

38

For a material to function as a suitable reference electrode there should be a stable

reversible reaction occurring on the surface of the conductor. Platinum is considered a pseudo

reference electrode that in these experiments turned out not to be very stable.

ESEM (environmental scanning electron microscope, Philips XL30 in the Earth and

Engineering Science building) images of the device were taken. Figure 4-9 shows the overall

image of the device, at the lowest magnification possible. The particles spotting the surface may

be dust particles. Figure 4-10 shows some of the corner and edges of the device that were imaged,

along with the location of the specific images. This was an untested device.

Figure 4-9: SEM image of the redesigned microelectronic device, at 20x magnification

Page 48: BETH A BIMBER - ETDA

39

These ESEM images show that there is either misalignment in a mask or the lift off

process (mask 2) removed too much titanium (working electrode). The end result is that platinum

(reference electrode) was showing. This ultimately means that the corrosion results from these

devices are not valid. Specific results from titanium implantable microelectronic devices will

have to wait until another round of devices is manufactured. Careful SEM images will also have

Figure 4-10: ESEM images of corners and edges of a titanium working electrode device sample, with location of the image shown on the device by the red circle

Page 49: BETH A BIMBER - ETDA

40

to be taken throughout the process, to insure cleanliness between the layers and on the surface of

the device. Another round of the second generation devices will have to be manufactured before

additional SEM images can be taken.

Magnesium Alloy Samples for In vitro testing

Corrosion test samples were prepared from wafers 144, 145, and 146 for in vitro testing

on the specific magnesium alloys. Open circuit potentials, along with the corrosion rate data from

both polarization resistance and EIS experiments were gathered. The EIS experimental results are

shown in the Bode format, which shows the response of the system to the changing frequency.

However, the Rp value was calculated from the Nyquist plot, as shown in Figure 2-8. The

experiments were run using a saturated calomel reference electrode.

Figure 4-11 show the results for sample 144-3 (a Mg 2.84 wt% Ti alloy). The open circuit

potential leveled off at a potential of -1.520 volts after about 200 seconds. The polarization

resistance curve did not pass through zero. Linear extrapolation of the curve provides a

polarization resistance value of 2518 Ω/cm2, which correlates to a corrosion rate of 4.14 mpy. The

EIS experiment provided a polarization resistance value of 2918.31 Ω/cm2, which correlates to a

corrosion value of 3.50 mpy that is in excellent agreement with the value obtained via the

polarization resistance method. Figure 4-12, shows the electrochemical results for sample 144-4

(another specimen taken from the Mg 2.84 wt% Ti alloy wafer), where again the curve does not

pass through zero. Linear extrapolation produces a polarization resistance of 1381 Ω/cm2. The

EIS experiment provides a polarization resistance value of 1479.9 Ω/cm2. While these

polarization resistance values result in higher corrosion rates than those observed on the other

specimen taken from this wafer, the values are still low for a magnesium alloy and within the

range of values needed for a bioasborable implant alloy.

Page 50: BETH A BIMBER - ETDA

41

Figure 4-13 shows the results for sample 145-4 (a Mg 3.97 wt% Ti alloy). The open

circuit potential levels off around 150 seconds at a potential of -1.476 volts, though after a drop in

the potential. The polarization resistance curve passes through zero and the polarization resistance

is calculated to be 217.46 Ω/cm2, while the EIS experiment produces a value of 2487.1 Ω/cm2.

Figure 4-14, the results from sample 145-6 (a Mg 3.97 wt% Ti alloy), has an open circuit

potential that levels off at a potential of -1.473 volts around 150 seconds. The polarization

resistance curve is linear and passes through zero; this provides a polarization resistance value of

294.47 Ω/cm2. A polarization resistance value of 1278.26 Ω/cm2 is produced from the EIS

experiment. Figure 4-15 shows the results for sample 145-7 (a Mg 3.97 wt% Ti alloy). The open

circuit potential appears to level off around 250 seconds at a potential of -1.466 volts. The

polarization resistance experiment provides a value of 175.16 Ω/cm2. The EIS experiment

provides a value of 182.96 Ω/cm2 for the polarization resistance. While there is a difference

between the rates given via EIS and polarization resistance, Petrilli also saw a lower corrosion

rate with the EIS technique [23]. The rates for the three polarization resistance experiments are

fairly similar, though slightly higher than the rates obtained from the magnesium 2.84 wt%

titanium samples.

Figure 4-16 shows the results for sample 146-4 (a Mg 7.44 wt% Ti alloy). The open

circuit potential leveled off to a potential of -1.507 volts around 220 seconds. The current density

passed through zero and the polarization resistance was calculated to be 60.96 Ω/cm2. The

polarization resistance value for the EIS experiment was calculated to be 182.96 Ω/cm2.

Page 51: BETH A BIMBER - ETDA

42

Figure 4-11: Open circuit potential, polarization resistance, and EIS results for sample 144-3 (Mg 2.84 wt% Ti) in HBSS at 37°C

-1.80

-1.75

-1.70

-1.65

-1.60

-1.55

-1.500 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.57

-1.56

-1.56

-1.55

-1.55

-1.54

-1.54

-1.53-2.5E-05 -2.0E-05 -1.5E-05 -1.0E-05 -5.0E-06

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-60

-50

-40

-30

-20

-10

0

10

20

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 52: BETH A BIMBER - ETDA

43

Figure 4-12: Open circuit potential, polarization resistance, and EIS results for sample 144-4 (Mg 2.84 wt% Ti) in HBSS at 37°C

-1.80

-1.75

-1.70

-1.65

-1.60

-1.55

-1.500 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.57

-1.56

-1.56

-1.55

-1.55

-1.54

-1.54

-1.53

-1.53-3.5E-05 -3.0E-05 -2.5E-05 -2.0E-05

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-60

-50

-40

-30

-20

-10

0

10

20

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 53: BETH A BIMBER - ETDA

44

Figure 4-13: Open circuit potential, polarization resistance, and EIS results for sample 145-4 (Mg 3.97 wt% Ti) in HBSS at 37°C

-1.66-1.64-1.62-1.60-1.58-1.56-1.54-1.52-1.50-1.48-1.46-1.44

0 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.50

-1.49

-1.49

-1.48

-1.48

-1.47

-1.47

-1.46

-1.46-2.0E-05 -1.0E-05 0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-60

-40

-20

0

20

40

60

80

100

120

140

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 54: BETH A BIMBER - ETDA

45

Figure 4-14: Open circuit potential, polarization resistance, and EIS results for sample 145-6 (Mg 3.97 wt% Ti) in HBSS at 37°C

-1.64-1.62-1.60-1.58-1.56-1.54-1.52-1.50-1.48-1.46-1.44

0 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.52

-1.51

-1.51

-1.50

-1.50

-1.49

-1.49

-1.48

-1.48-3.0E-05 -2.0E-05 -1.0E-05 0.0E+00 1.0E-05 2.0E-05

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-50

-40

-30

-20

-10

0

10

20

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 55: BETH A BIMBER - ETDA

46

Figure 4-15: Open circuit potential, polarization resistance, and EIS results for sample 145-7 (Mg 3.97 wt% Ti) in HBSS at 37°C

-1.70

-1.65

-1.60

-1.55

-1.50

-1.450 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.49

-1.49

-1.48

-1.48

-1.47

-1.47

-1.46

-1.46

-1.45-5.0E-05 -3.0E-05 -1.0E-05 1.0E-05 3.0E-05

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-50-45-40-35-30-25-20-15-10-505

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 56: BETH A BIMBER - ETDA

47

Figure 4-16: Open circuit potential, polarization resistance, and EIS results for sample 146-4 (Mg 7.44 wt% Ti) in HBSS at 37°C

-1.75

-1.70

-1.65

-1.60

-1.55

-1.50

-1.450 100 200 300

OCP

vs S

CE (V

)

Time (seconds)

-1.46

-1.46

-1.45

-1.45

-1.44

-1.44

-1.43

-1.43

-1.42-6.0E-05 -4.0E-05 -2.0E-05 0.0E+00 2.0E-05 4.0E-05 6.0E-05

Pote

ntia

l vs

SCE

(V)

Current Density (A/cm2)

-50

-40

-30

-20

-10

0

10

20

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E-02 1.0E+00 1.0E+02 1.0E+04

Z phz

(°)

Z mod

(ohm

)

Frequency (Hz)

Page 57: BETH A BIMBER - ETDA

48

Sample Calculations for Corrosion Rate Determination

Equations 2-7 and 2-8 were used to calculate the corrosion rate of all of the samples. An

example calculation for sample 144-3 (Mg 2.84 wt% Ti alloy) is presented in this section. The

Tafel slopes, ba and bc used were 0.100 volts/decade. The polarization resistance, Rp is the slope

of the potential versus current density where the curve passes through zero.

𝑖𝐶𝑂𝑅𝑅 = 2.303 � 𝑏𝑎𝑏𝑐𝑏𝑎+𝑏𝑐

� � 1𝑅𝑝� = 2.303 �0.1∗0.1

0.1+0.1� � 1

0.0252� = 4.572 𝜇𝐴

𝑐𝑚2

The corrosion rate can then be calculated. The constant K is 0.129. The density is 1.738 g/cm3

and the equivalent weight is 12.2 g.

𝐶𝑜𝑟𝑟𝑜𝑠𝑖𝑜𝑛 𝑅𝑎𝑡𝑒 = 𝐾 𝑖𝐶𝑂𝑅𝑅𝜌

𝐸𝑊 = 0.1294.572 𝜇𝐴𝑐𝑚2

1.738 𝑔𝑐𝑚3

(12.2𝑔) = 4.14 𝑚𝑝𝑦

Mils per year can be converted into mm/day:

𝑚𝑚𝑑𝑎𝑦

= �0.0254𝑚𝑚

𝑦𝑟

1 𝑚𝑝𝑦� ∗ (# 𝑚𝑝𝑦) ∗ � 1 𝑦𝑟

365 𝑑𝑎𝑦𝑠�

Comparison of Corrosion Rates and Examination of Surface Morphology

Table 4-3 summarizes the corrosion rates obtained from the data shown in Figures 4-11

through 4-16. The open circuit potential was run for an additional 30 seconds between the

polarization resistance and EIS experiments.

Equation 4-1

Equation 4-2

Equation 4-3

Page 58: BETH A BIMBER - ETDA

49

As Table 4-3 reveals, the corrosion rate is dependent on the composition of the magnesium-

titanium alloy. A composition around Mg 2.84 wt% Ti produces the lowest corrosion rates. These

rates (of a couple of mils per year (or on the order of 10-4 mm/day) are comparable to those

obtained by Petrilli [23] for e-beam evaporated magnesium alloys with approximately 2%

titanium. These electrochemical results reveal that the e-beam evaporated magnesium-2.84%

titanium alloy has the desired corrosion rate for a bioabsorable implantable device. However,

some processing issues with the three electrode device still need to be overcome before the

devices will pass in vitro testing and the overall research program can continue into the in vivo

animal testing stage.

SEM images were also taken, prior to electrochemical testing, to examine the

morphology of the magnesium-titanium alloy. Figure 4-17 shows the SEM images of sample

145-3 (Mg 3.97 wt% Ti alloy).

Table 4-3: Open circuit potentials and corrosion rates (via polarization resistance and EIS experiments) of various magnesium samples versus SCE (reference) in HBSS at 37°C

Mg sample versus SCE

Sample Composition PR EIS # wt% Ti OCP mpy mm/day Rp (Ω) mpy mm/day

144-3 2.84 -1.549 4.14 2.88E-4 2708.3 3.85 2.68E-4 144-4 2.84 -1.545 7.55 5.25E-4 786.95 13.25 9.22E-4 145-4 3.97 -1.477 47.95 33.4E-4 2487.1 4.19 2.92E-4 145-6 3.97 -1.500 35.41 24.6E-4 1278.3 8.15 5.68E-4 145-7 3.97 -1.472 59.53 41.4E-4 2577.7 4.04 2.81E-4 146-4 7.44 -1.443 171.04 119E-4 182.96 56.99 39.7E-4

Page 59: BETH A BIMBER - ETDA

50

This image shows what looks like filiform corrosion – a special form of localized crevice

corrosion that occurs under a thin oxide film. It can be recognized by emancipating cracks, which

are fine tunnels worming just beneath the surface [38]. Figure 4-18 shows that this corrosion

product is all over the surface of this particular sample.

Figure 4-17: ESEM image of the top of the magnesium sample 145-3 (Mg 3.97 wt% Ti alloy) prior to electrochemical testing

Page 60: BETH A BIMBER - ETDA

51

However, filiform corrosion is a form of surface corrosion, impacting just beneath the surface

layer. It does not burrow into the entire sample and render the alloy useless. This type of attack

while cosmetically unattractive is relatively harmless (as the corrosion rates in Table 4-3 reveal)

and indicated that a more protective oxide layer is formed on the Mg-Ti alloy than on typical

magnesium. More samples need to be run, after more samples have been created.

Figure 4-18: ESEM image of the top of the magnesium sample 145-3 (Mg 3.97 wt% Ti alloy), at a lower magnification, prior to electrochemical testing

Page 61: BETH A BIMBER - ETDA

52

Chapter 5

Conclusions

The use of magnesium as a biodegradable medical implant is viable as the corrosion rates

in Table 4-3 reveal. In vitro testing of the various magnesium-titanium alloys is still underway.

The results of this senior thesis have shown that a magnesium 2.84% titanium alloy has adequate

corrosion resistance to proceed into final in vitro testing, followed by in vivo testing, once a few

small processing issues in the device are fixed.

The microelectronic implantable device can be implanted and utilized as a means of

measuring the corrosion rate. At this current point in time, more in vitro testing needs to be

conducted to insure that the microelectronic device is functioning properly. The specific in vivo

corrosion data can provide valuable insight into the behavior of specific magnesium alloys in the

host body, which can then provide information needed to one day construct biodegradable stents

and bone plates.

Page 62: BETH A BIMBER - ETDA

53

Chapter 6

Future Work

The next step to testing the microelectronic implantable devices is to conduct more in

vitro tests. The experimentally measured corrosion rate of the working electrode of the device

needs to match the corrosion rates of the samples versus sputtered platinum as the counter and

reference electrodes. Once the device proves to be in working condition, then the issue of

connections can be further addressed.

Wire connections will be attached to the electrical pads on the device. However, the wires

are very thin; sturdier wires need to connect to these tiny wires. The studier wires will then be the

wires that emerge from the rat’s skull during the in vivo corrosion experiments. Figure 6-1 shows

the positioning of the device in the rat’s skull and a cap similar to the one that will be used, on a

rat’s head. This cap protects any protruding wires and the incision site from the claws of the rats.

Figure 6-1: Image of a rat’s skull, with the implant location of the microelectronic implantable device (left), Skull protection cap on a rat

Page 63: BETH A BIMBER - ETDA

54

Current research is also being conducted on the sterilization practices of other implants.

Before insertion, this microelectronic device needs to be deemed sanitary enough for

implantation, based on protocols from the Institutional Animal Care and Use Committee.

The next step is implantation and the running of the corrosion experiments. Open circuit

potential and polarization resistance experiments will be performed. From these experiments, in

vivo corrosion rates can be obtained. These results will provide knowledge into the difference

between in vitro and in vivo corrosion rates, and provide insight into which specific magnesium-

titanium alloy would be best for a biodegradable medical implant. Further projection of this

research leans towards coronary stents and bone plates.

Page 64: BETH A BIMBER - ETDA

55

Appendix A

Simulated Body Fluids Recipe

HBSS Recipe [39]:

Table A-1: Recipe for HBSS, creates 1 L of solution

Component Concentration (mg/L) CaCl2·2H2O 186

KCl 400 KH2PO4 60 MgSO4 98 NaCl 8000

NaHCO3 350 Na2HPO4·7H2O 90

Glucose 1000

Page 65: BETH A BIMBER - ETDA

56

Appendix B

Detailed Recipe for Creating Devices

The procedure for creating an implantable microelectronic device was derived from a

master’s thesis titled “Study of Magnesium Corrosion for Biomedical Devices: An Investigation

Involving Microelectronic Processing” by Anna Hartsock. This thesis should be consulted for

more complete details. The procedure that is presented in this thesis is a paraphrased version of

what is contained in the thesis by Anna Hartsock.

Mask One

A photoresist assist (LOR2A) and a photoresist (S1813) are spun onto a six inch wafer.

The phototresist assist was spun for 45 seconds at 3000 rpm and baked at 180°C for five minutes.

The photoresist is spun at 4000 rpm for 45 seconds and baked at 95°C for one minute.

The mask size left large unused area; to make better use of the wafer area, the wafers

were cleaved into quarters. A Karl Suss MA BA6 mask aligner was used to align Mask One. The

wafer is then exposed (hard contact exposure) to UV light at 8 kW/cm2 for ten seconds. Exposure

to UV light will break the chemical bonds in the photoresist, allowing the exposed resist to be

removed easier. The developer to remove the photoresist is CD-26. The quarters were immersed

for about one minute, or until the exposed photoresist was completely removed. The wafers were

then rinsed in DI water and dried with nitrogen.

The wafers were loaded into the Kurt J. Lesker CMS System at the Materials Research

Lab (MRL). First, a cleaning recipe was used to remove any dust particles and phototresist. Then

titanium and platinum were sputtered onto the wafer. Titanium was sputtered for 120 seconds at

100Å and the platinum was sputtered at 1000Å for 400 seconds on top of the titanium layer. The

quarters were then immersed in an acetone bath and ultrasonicated, to remove the platinum and

Page 66: BETH A BIMBER - ETDA

57

titanium that was on top of the photoresist.

Mask Two

The quarters were then transferred to the EB-PVD system at the Earth-Engineering

Sciences Building. The working electrode was deposited on the wafer. Phototresist (S1813) was

spun on the quarter wafer for 45 seconds at 4000 rpm and baked at 110°C for one minute. Mask

Two was aligned and UV light was exposed to the system for seven seconds. The quarter was

placed in a developer bath (CD-26) for about one minute to remove the photoresist.

A bath of HNO3 :H2O (ratio of 20 to 180 mL) was then utilized to remove the exposed

working electrode. The quarter was immersed for roughly 50 seconds. A rinse of acetone and IPA

insured that all of the excess deposited working electrode was removed before the next step.

Mask Three

The quarter wafers were loaded into the deposition chamber of the parylene evaporator.

A recipe for evaporating parylene to obtain a desired thickness of 0.5μm was utilized. Actual

thickness, when tested later, determined that the thickness was not uniform, but was between

0.56μm and 0.67μm. Photoresist (S1813) was spun on at a rate of 4000 rpm for 45 seconds. UV

light was exposed for ten seconds and immersed in a developer bath (CD-26) for about one

minute.

An oxygen plasma etcher was used to remove the exposed parylene coating on the

electrode and the connection pads of the electrical leads. The PT720 plasma machine used an

oxygen plasma composed of 90% O2 and 10% Ar. The quarters were exposed in 30 second

increments until the exposed parylene was removed.

Page 67: BETH A BIMBER - ETDA

58

References

[1] Winter, M. (2011). Magnesium. Retrieved July 7, 2011, from WebElements:

http://www.webelements.com/magnesium/

[2] Moloi, Kagiso. (2011) The story of densitry: Ancient origins. Namibian Dental

Association. Retrieved July 7, 2011, from

http://www.namibiadent.com/History/HistoryDentistry.html

[3] Wood, J.L. (2011) Surgical Plates and Screws History. eHow Health. Retrieved July

7, 2011, from http://www.ehow.com/about_6707980_surgical-plates-screws-history.html

[4] (2011) Dietary Supplement Fact Sheet: Magnesium. Office of Dietary Supplements

National Institutes of Health. Retrieved July 8, 2011 from

http://ods.od.nih.gov/factsheets/magnesium/

[5] Xu, L., Pan, F., Yu, G., Yang, L., Zhang, E., & Yang, K. (2009). In vitro and in vivo

evaluation of the surface bioactivity of a calcium phosphate coated magnesium alloy.

Biomaterials, 30(8), 1512-23. Elsevier Ltd. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/19111896

[6] Zeng, R., Dietzel, W., Witte, F., Hort, N., & Blawert, C. (2008). Progress and

Challenge for Magnesium Alloys as Biomaterials. Advanced Engineering Materials, 10(8), B3-

B14. Retrieved from http://doi.wiley.com/10.1002/adem.200800035

Page 68: BETH A BIMBER - ETDA

59

[7] Witte, F., Fischer, J., Nellesen, J., Crostack, H.-A., Kaese, V., Pisch, A., Beckmann,

F., et al. (2006). In vitro and in vivo corrosion measurements of magnesium alloys. Biomaterials,

27(7), 1013-8. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/16122786

[8] Shaw, B. A. (2003). Corrosion Resistance of Magnesium Alloys. ASM Handbook,

Volume 13A Corrosion: Fundamentals, Testing and Protection (Vol. 13, pp. 692-696).

[9] Heublein, B., Rohde, R., Kaese, V., Niemeyer, M., Hartung, W., & Haverich, A.

(2003). Biocorrosion of magnesium alloys: a new principle in cardiovascular implant technology?

Heart (British Cardiac Society), 89(6), 651-6. Retrieved from

http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1767674&tool=pmcentrez&rendertyp

e=abstract

[10] Xu, L., Yu, G., Zhang, E., Pan, F., & Yang, K. (2007). In vivo corrosion behavior of

Mg-Mn-Zn alloy for bone implant application. Journal of Biomedical Materials Research Part A,

83(3), 703-711.

[11] Hort N, Huang Y, Fechner D, et al. Magnesium alloys as implant materials--

principles of property design for Mg-RE alloys. Acta biomaterialia. 2010;6(5):1714-25. Available

at: http://www.ncbi.nlm.nih.gov/pubmed/19788945 [Accessed July 27, 2011].

[12] Zhang, E., Xu, L., Yu, G., Pan, F. and Yang, K. (2009), In vivo evaluation of

biodegradable magnesium alloy bone implant in the first 6 months implantation. Journal of

Biomedical Materials Research Part A, 90A: 882–893. doi: 10.1002/jbm.a.3213

Page 69: BETH A BIMBER - ETDA

60

[13] Witte, F. (2005). In vivo corrosion of four magnesium alloys and the associated bone

response. Biomaterials, 26(17), 3557–63. Retrieved from

http://www.sciencedirect.com/science/article/B6TWB-4DPC673-

4/2/0a166382be2626fc15597610f6fdd974

[14] Witte, F., Fischer, J., Nellesen, J., Vogt, C., Vogt, J., Donath, T., & Beckmann, F.

(2010). In vivo corrosion and corrosion protection of magnesium alloy LAE442. Acta

Biomaterialia, 6(5), 1792-1799. Acta Materialia Inc. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/19822226

[15] Witte F, Fischer J, Nellesen J, Beckmann F. Microtomography of magnesium

implants in bone and their degradation. Proceedings of SPIE. 2006;6318(Table 1):631806-

631806-9. Available at: http://link.aip.org/link/PSISDG/v6318/i1/p631806/s1&Agg=doi

[Accessed June 29, 2011].

[16] Donner, S., Witte, F., Bartsch, I., Petraglia, F., Massow, O., Heidrich, M., et al.

(2010). In vivo optical coherence tomography of percutaneous implants in hairless mice.

Biomedicine, 7554, 75540S-75540S-6. doi: 10.1117/12.841640.

[17] Witte F. The history of biodegradable magnesium implants: a review. Acta

biomaterialia. 2010;6(5):1680-92. Available at: http://www.ncbi.nlm.nih.gov/pubmed/20172057

[Accessed July 27, 2011].

Page 70: BETH A BIMBER - ETDA

61

[18] Witte F, Eliezer A, Cohen S. The History, Challenges and the Future of

Biodegradable Metal Implants. Advanced Materials Research. 2010;95:3-7. Available at:

http://www.scientific.net/AMR.95.3 [Accessed September 29, 2011].

[19] Eliezer A, Witte F. Corrosion Behavior of Magnesium Alloys in Biomedical

Environments. Advanced Materials Research. 2010;95:17-20. Available at:

http://www.scientific.net/AMR.95.17 [Accessed September 29, 2011].

[20] Zeng R, Dietzel W, Witte F, Hort N, Blawert C. Progress and Challenge for

Magnesium Alloys as Biomaterials. Advanced Engineering Materials. 2008;10(8):B3-B14.

Available at: http://doi.wiley.com/10.1002/adem.200800035 [Accessed June 19, 2011].

[21] Witte F, Fischer J, Nellesen J, Beckmann F. Microtomography of magnesium

implants in bone and their degradation. Proceedings of SPIE. 2006;6318(Table 1):631806-

631806-9. Available at: http://link.aip.org/link/PSISDG/v6318/i1/p631806/s1&Agg=doi

[Accessed June 29, 2011].

[22] Wolfe, R.C. (2005) Imparting Passivity to Vapor Deposited Magnesium Alloys.

Doctoral dissertation. The Pennsylvania State University, University Park, PA.

[23] Petrilli, J.D. (2009) Evaluating the In Vitro Corrosion Behavior and Cytotoxicity of

Vapor Deposited Magnesium Alloys. Master’s thesis. The Pennsylvania State University,

University Park, PA.

Page 71: BETH A BIMBER - ETDA

62

[24] Van Pelt, Jacqueline. (2009) Magnesium Alloys For Use in Bioabsorable Cardiac

Stents. Bachelor’s thesis. The Pennsylvania State University. University Park, PA.

[25] Pytel, Nicholas. (2009) Implant design for real time in vivo analysis of a novel

bioabsorbable magnesium alloy. Bachelor’s thesis. The Pennsylvania State University, University

Park, PA.

[26] Winter, M. (2011) Titanium. WebElements. Retrieved July 9, 2011 from

http://www.webelements.com/titanium/

[27] Winter, M. (2011) Aluminum. WebElements. Retrieved July 9, 2011 from

http://www.webelements.com/aluminum/

[28] Winter, M. (2011) Yttrium. WebElements. Retrieved July 9, 2011 from

http://www.webelements.com/yttrium/

[29] Feyerabend F, Fischer J, Holtz J, et al. Evaluation of short-term effects of rare earth

and other elements used in magnesium alloys on primary cells and cell lines. Acta biomaterialia.

2010;6(5):1834-42. Available at: http://www.ncbi.nlm.nih.gov/pubmed/19800429 [Accessed July

27, 2011].

[30] (1994). Surface Coating by Physical Vapor Deposition. ASM Handbook, Volume

5Surface Engineering Retrieved from

http://205.153.241.230/P2_Opportunity_Handbook/1_5.html

Page 72: BETH A BIMBER - ETDA

63

[31] (2011) Linear Polarization Resistance. Corrosion Doctors. Retrieved September 3,

2011 from http://corrosion-doctors.org/Electrochem/LPR.htm

[32] Jones, Denny A. Principles and Prevention of Corrosion. 2nd ed. Upper Saddle

River: Pearson plc, 1996.

[33] Silverman, D.C. (2011) Tutorial on Corrosion Rate Calculator Argetum Solutions,

Inc.. Retrieved September 21, 2011 from

http://www.argentumsolutions.com/tutorials/calculator_tutorialpg4.html

[34] (2011) Corrosion Rate Calculator. Corrosion Source. Retrieved September 21, 2011

from http://www.corrosionsource.com/Corrosion-Rate-Calculator

[35] Hamdy, A. S. (2006). Electrochemical Impedance Spectroscopy Study of the

Corrosion Behavior of Some Niobium Bearing Stainless Steels. International Journal of

Electrochemical Science, 1, 171-180.

[36] Gamry Instruments: Basics of Electrochemical Impedance Spectroscopy. (2011).

Retrieved from http://www.gamry.com/App_Notes/EIS_Primer/Basics_Of_ EIS.pdf.

[37] Haro, C. de, Mas, R., Abadal, G., Muñoz, J., Perez-Murano, F., & Dominguez, C.

(2002). Electrochemical platinum coatings for improving performance of implantable

microelectrode arrays. Biomaterials, 23(23), 4515-21. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/12322971.

Page 73: BETH A BIMBER - ETDA

64

[38] (2011). Different Types of Corrosion: Filiform Corrosion (Underfilm Corrosion).

WebCorr Corrosion Consulting Services. Retrieved October 30, 2011 from

http://www.corrosionclinic.com/types_of_corrosion/filiform_corrosion_underfilm_corros

ion.htm

[39] Hartsock, Anna. (2012) Study of Magnesium Corrosion for Biomedical

Devices: An Investigation Involving Microelectronic Processing. Unpublished Master's

Thesis. The Pennsylvania State University. University Park, PA.

Page 74: BETH A BIMBER - ETDA

65

VITA

Beth Bimber

College Address 127 E. Hamilton Ave, Apt 8, State College, PA 16801 Home Address 5568 Coachmans Lane, Hamburg, NY 14075 Phone (716) 868-2112 Citizenship USA Educational History The Pennsylvania State University, University Park, PA Major Engineering Science, the Honors Engineering Curriculum Minor Engineering Mechanics Degree B.S. 2011 Honors The Schreyer Honors College Scholar 2008-2011 Honors Thesis Implantable Microelectronic Devices for Measuring In vivo

Corrosion of Magnesium Alloys (Thesis Supervisor: Dr. Barbara Shaw) Professional Positions Drafter, LORD Corporation, Erie, PA. Full-time summer position. 2009

Duties: Supervisor: Student Researcher, University of Houston, Houston, TX. Full-time summer position. 2010

Duties: Supervisor:

Honors Student Intern, Drivetrain Center at the Applied Research Lab, University Park, PA.

Full-time summer position, half-time fall position. 2011 Duties: Supervisor:

Student Researcher, Corrosion Lab, University Park, PA. Part-time unpaid position. 2011-present

Duties: Supervisor:

Professional Associations Golden Key International Honor Society (student member) Phi Kappa Phi Honor Society (student member) Omnicrom Delta Kappa Honor Society (student member) Professional Activities Secretary and Webmaster, Society of Engineering Science Pennsylvania State University chapter, 2011.

Rube Goldberg Team Member, American Society of Mechanical Engineers Pennsylvania State University chapter, 2011.