+ All Categories
Home > Documents > Passivation of aluminum with alkyl phosphonic acids for biochip...

Passivation of aluminum with alkyl phosphonic acids for biochip...

Date post: 17-Mar-2021
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
5
Applied Surface Science 256 (2010) 7146–7150 Contents lists available at ScienceDirect Applied Surface Science journal homepage: www.elsevier.com/locate/apsusc Passivation of aluminum with alkyl phosphonic acids for biochip applications Sachin Attavar a , Mohit Diwekar a , Matthew R. Linford b , Mark A. Davis c , Steve Blair a,a Department of Electrical Engineering, University of Utah, 50 South Central Campus Drive, Salt Lake City, UT 84112, United States b Department of Chemistry and Biochemistry, Brigham Young University, Provo, UT 84602, United States c MOXTEK Inc, 452 West 1260 North, Orem, UT 84057, United States article info Article history: Received 16 March 2010 Received in revised form 11 May 2010 Accepted 11 May 2010 Available online 19 May 2010 Keywords: Biosenor DNA metallic microarray Phosphonic acid self-assembled monolayer abstract Self-assembly of decylphosphonic acid (DPA) and octadecylphosphonic acid (ODPA) was studied on alu- minum films using XPS, ToF-SIMS and surface wettability. Modified aluminum films were tested for passivation against silanization and subsequent oligonucleotide attachment. Passivation ratios of at least 450:1 compared to unprotected aluminum were obtained, as quantified by attachment of radio-labeled oligos. © 2010 Elsevier B.V. All rights reserved. 1. Introduction During the last two decades, the use of self-assembled mono- layers (SAM’s) in the production of well-defined surfaces has undergone tremendous growth, advanced in part through the depth of surface characterization methods [1,2]. SAM’s are ordered molecular assemblies that are spontaneously formed by the adsorption of molecules with head groups that show affinity to a specific substrate, which enables ordering on the surface without any preassembly [2]. One of the most active areas of research with SAM’s is towards the production of biochips. The fabrication and manipulation of molecular assemblies, combined with molecular recognition and computational chemistry to elucidate structure–function rela- tionships, forms the core of modern surface chemistry. One of the important design aspects for biochips is to isolate probing molecules into active areas of transduction. This paper dis- cusses the development of a biosensor array for ultimate use in molecular diagnostics. The role of SAM’s to form chemically dis- tinct regions on the patterned aluminum surface of a biochip is explored. In the literature, alkanethiols on gold [3–5] and alkylsilanes on silica [6] are the two most extensively studied classes of SAM’s. Recent literature has also seen an array of functionalities devel- oped that can be used with more active metals. Studies have been reported for titanium [7–10], tantalum [11–14] and aluminum [15,16] oxides coated with alkylsilanes, thiols, phosphonates, car- Corresponding author. E-mail address: [email protected] (S. Blair). boxylic acids, and so forth. In this paper, the aluminum oxide and phosphonic acid system will be explored. Aluminum, being a light metal, bears potential in a broad range of technological applica- tions. Aluminum can be classified as an active metal due to its tendency to oxidize quickly. The aluminum oxide layer is chem- ically bound to the surface, and it seals the core aluminum from any further reaction. Aluminum has been used in the development of biosensors due to its dielectric properties supporting, for example, the propagation and localization of surface plasmons. These biosensors are fabri- cated by patterning an aluminum layer on top of a glass/quartz substrate, where the pattern often takes the form of microscopic holes [17,18]. An aluminum layer can also be patterned macro- scopically by opening “windows” to the underlying substrate that precisely define the location and morphology of capture zones [19]. For DNA-based biosensors on glass, the primary anchoring molecules are silanes. For development of a similar biochip with aluminum as the metallic component, it is desirable to prevent attachment of molecular recognition functionality (silane) to the metal film. A selective and localized passivation approach is needed for these kinds of mixed material biochips. The most commonly used sensors are based upon glass in conjunction with gold. Bare alu- minum poses a problem as silane attaches to both aluminum (AL–O–Si) and glass (Si–O–Si). Hence prior to silanization, alu- minum needs to be passivated. Alkyl phosphonic acids are reported to passivate a variety of metal oxides, such as titanium or aluminum oxide, while not binding to SiO 2 surfaces in an aqueous medium [18,20,21]. This represents a potential approach for targeted molec- ular placement of capture oligos on patterned aluminum film biochips. 0169-4332/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2010.05.041
Transcript
Page 1: Passivation of aluminum with alkyl phosphonic acids for biochip …test5.coe.utah.edu/files/2014/07/Passivation-of-aluminum.pdf · 2014. 8. 1. · It is known that the phosphonic

Applied Surface Science 256 (2010) 7146–7150

Contents lists available at ScienceDirect

Applied Surface Science

journa l homepage: www.e lsev ier .com/ locate /apsusc

Passivation of aluminum with alkyl phosphonic acids for biochip applications

Sachin Attavara, Mohit Diwekara, Matthew R. Linfordb, Mark A. Davisc, Steve Blaira,∗

a Department of Electrical Engineering, University of Utah, 50 South Central Campus Drive, Salt Lake City, UT 84112, United Statesb Department of Chemistry and Biochemistry, Brigham Young University, Provo, UT 84602, United Statesc MOXTEK Inc, 452 West 1260 North, Orem, UT 84057, United States

a r t i c l e i n f o

Article history:Received 16 March 2010Received in revised form 11 May 2010Accepted 11 May 2010Available online 19 May 2010

Keywords:BiosenorDNA metallic microarrayPhosphonic acid self-assembled monolayer

a b s t r a c t

Self-assembly of decylphosphonic acid (DPA) and octadecylphosphonic acid (ODPA) was studied on alu-minum films using XPS, ToF-SIMS and surface wettability. Modified aluminum films were tested forpassivation against silanization and subsequent oligonucleotide attachment. Passivation ratios of at least450:1 compared to unprotected aluminum were obtained, as quantified by attachment of radio-labeledoligos.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

During the last two decades, the use of self-assembled mono-layers (SAM’s) in the production of well-defined surfaces hasundergone tremendous growth, advanced in part through thedepth of surface characterization methods [1,2]. SAM’s are orderedmolecular assemblies that are spontaneously formed by theadsorption of molecules with head groups that show affinity to aspecific substrate, which enables ordering on the surface withoutany preassembly [2].

One of the most active areas of research with SAM’s is towardsthe production of biochips. The fabrication and manipulationof molecular assemblies, combined with molecular recognitionand computational chemistry to elucidate structure–function rela-tionships, forms the core of modern surface chemistry. One ofthe important design aspects for biochips is to isolate probingmolecules into active areas of transduction. This paper dis-cusses the development of a biosensor array for ultimate use inmolecular diagnostics. The role of SAM’s to form chemically dis-tinct regions on the patterned aluminum surface of a biochip isexplored.

In the literature, alkanethiols on gold [3–5] and alkylsilanes onsilica [6] are the two most extensively studied classes of SAM’s.Recent literature has also seen an array of functionalities devel-oped that can be used with more active metals. Studies have beenreported for titanium [7–10], tantalum [11–14] and aluminum[15,16] oxides coated with alkylsilanes, thiols, phosphonates, car-

∗ Corresponding author.E-mail address: [email protected] (S. Blair).

boxylic acids, and so forth. In this paper, the aluminum oxide andphosphonic acid system will be explored. Aluminum, being a lightmetal, bears potential in a broad range of technological applica-tions. Aluminum can be classified as an active metal due to itstendency to oxidize quickly. The aluminum oxide layer is chem-ically bound to the surface, and it seals the core aluminum fromany further reaction.

Aluminum has been used in the development of biosensors dueto its dielectric properties supporting, for example, the propagationand localization of surface plasmons. These biosensors are fabri-cated by patterning an aluminum layer on top of a glass/quartzsubstrate, where the pattern often takes the form of microscopicholes [17,18]. An aluminum layer can also be patterned macro-scopically by opening “windows” to the underlying substrate thatprecisely define the location and morphology of capture zones[19]. For DNA-based biosensors on glass, the primary anchoringmolecules are silanes. For development of a similar biochip withaluminum as the metallic component, it is desirable to preventattachment of molecular recognition functionality (silane) to themetal film.

A selective and localized passivation approach is needed forthese kinds of mixed material biochips. The most commonly usedsensors are based upon glass in conjunction with gold. Bare alu-minum poses a problem as silane attaches to both aluminum(AL–O–Si) and glass (Si–O–Si). Hence prior to silanization, alu-minum needs to be passivated. Alkyl phosphonic acids are reportedto passivate a variety of metal oxides, such as titanium or aluminumoxide, while not binding to SiO2 surfaces in an aqueous medium[18,20,21]. This represents a potential approach for targeted molec-ular placement of capture oligos on patterned aluminum filmbiochips.

0169-4332/$ – see front matter © 2010 Elsevier B.V. All rights reserved.doi:10.1016/j.apsusc.2010.05.041

Page 2: Passivation of aluminum with alkyl phosphonic acids for biochip …test5.coe.utah.edu/files/2014/07/Passivation-of-aluminum.pdf · 2014. 8. 1. · It is known that the phosphonic

S. Attavar et al. / Applied Surface Science 256 (2010) 7146–7150 7147

One of the main reasons for using phosphonic acids ratherthan the well-known carboxylic acids is their stronger bindingwith the oxide [22,23]. Aluminum forms a native oxide whenexposed to an oxygen-containing environment. The phospho-nic acids are mostly deposited from an organic or water-diluted(10−3 mol/l) solution [1]. It is known that the phosphonic acidsinteract with the aluminum hydroxyl groups, where an increasein the number of surface hydroxyls enhances the phospho-nic acid deposition [24,25]. Phosphonic acid specifically reactsto hydrated aluminum oxide, through the Al–O–P bond. TheSi–O–P bond formed on glass substrates is easily hydrolysable[20]. The phosphonic acid prevents subsequent chemical treat-ments, such as exposure to silane containing molecules, fromreacting with the aluminum. Further, capture molecules can beattached to non-aluminum surfaces via reaction with a specificfunctional group of the silane. However, tailoring a surface throughthe attachment of phosphonic acid SAM’s requires careful alu-minum surface preparation prior to monolayer self-assembly [24].The main issue related to the use of the aluminum–phosphonicacid system is the question of whether a dense SAM can beformed. Although for many systems it has been proven thatthe chain length is a crucial factor in the formation of a SAM[23], it has been shown that this is not the case for phosphonicacids [26,27]. Relatively well-oriented and stable films have beenformed using pentaphosphonic acid [26] and octylphosphonic acid[1].

In this paper, we describe the preparation and characterizationof self-assembled monolayers of alkylphosphonic acids (primar-ily octadecyl phosphonic acid and decyl phosphonic acid) on analuminum oxide surface. As described previously, this method hasbeen extensively used for purposes of corrosion protection. Thispaper explores the potential of the alkyl phosphonic acid layer to beused as an agent to improve selective silanization of sensors basedon metallic array structures. The other aspect of this study will be toinvestigate the role played by alkyl chain length of phosphonic acidin formation of an effective passivation layers towards silanization.This study further characterizes passivation against immobiliza-tion of oligonucelotide molecules as a step towards developmentof DNA-based diagnostic biochips.

2. Experimental

2.1. Chemicals

All chemicals are reagent grade. The phosphonic acidsemployed were n-butylphosphonic acid—CH3(CH2)3PO(OH)2(BPA), n-decylphosphonic acid—CH3(CH2)9PO(OH)2 (DPA) andn-octadecylphosphonic acid—CH3(CH2)17PO(OH)2 (ODPA).They were purchased from Alfa Aesar (purity 98%). 3-Glycidoxypropyltrimethoxysilane—CH2OCHCH2O(CH2)3Si(OCH3)3(GPS) was purchased from Sigma–Aldrich. A terminal transferaselabeling kit was purchased from New England Biolabs. [!-32P]dATP was purchased from Perkin Elmer.

2.2. Substrate preparation

Moxtek Inc. (Provo, UT) provided plain glass substrates andsubstrates coated with 100 nm aluminum films. Windows in Alwere fabricated with conventional photolithography: the Al sub-strates were coated with a positive photoresist then soft baked.A photomask with an array of 200 "m × 200 "m square holeswith a periodicity of 400 "m was generated using an Electro-mask MM250 pattern generator. An Electronic Visions EV-420mask aligner was used to transfer the pattern onto the photore-sist. After developing the photoresist, the exposed Al areas were

etched out with a standard Al-etchant and the photoresist wasremoved.

The substrates were cleaned using solvent wash. The washincluded acetone, isopropyl alcohol and methanol. After solventwash, samples were rinsed with doubly distilled water (ddH2O) anddried using nitrogen. This was followed by oxygen plasma cleaningusing a Harrick plasma cleaner. The plasma cleaner was operatedat the medium power setting (200 W) for 5 min. At this point, thesamples were placed in a closed petri dish and set aside for 30 minbefore any further processing.

2.3. Self-assembly

Phosphonic acid solutions of 1 mM were prepared in methanol,a concentration that is normally reported in the literature as themolecules behave as free species in the solution. The passivationlayer was self-assembled onto the substrates by soaking them inphosphonic acid solution at room temperature for 48 h. Sampleswere cleaned in methanol and dried under nitrogen. After passi-vation, samples were annealed under nitrogen for 4 h at 90 ◦C. Thephysisorbed phosphonic acid was removed using triple methanoland water washes.

2.4. Silanization

After cleaning, the substrates were placed in a Fisher Sci-entific oven at 115 ◦C with a small vial containing 1.5 ml of3-glycidoxypropyltrimethoxysilane (GPS). The oven was sealed,pumped down, and purged three times with ultrapure nitrogen.After 8 h, the oven was purged with nitrogen and the substrateswere removed.

2.5. Surface characterization

Surface wettability was investigated by measuring the contactangle in a sessile water drop experiment. A water drop of 1 ml vol-ume was used in each measurement. Three independent readingswere taken for each sample.

The thicknesses were measured by reflectance variable anglespectroscopic ellipsometry (VASE) using material databases forunderlying films (Al on glass, Al2O3). The measurements were car-ried out in the spectral range of 300–1000 nm. The measured ! and" for the phosphonic acid layer were fitted using “thickness” as thefitting parameter. The fitting procedure was repeated for measure-ments at various positions on the sample surface to get the averagethickness for each film.

XPS analyses were performed on an Axis Ultra spectrometerfrom Kratos (Manchester, U.K.) equipped with a concentric hemi-spherical analyzer and using a mono-chromatized aluminum anodeX-ray source maintained at 15 KeV. The samples were investigatedunder ultrahigh vacuum conditions: 10−8 to 10−7 Pa. Samples wereanalyzed with a pass energy of 160.0 eV for survey scans and 40.0 eVfor high energy resolution elemental scans.

Static ToF-SIMS (Cameca/ION-TOF IV SIMS) was performed witha monoisotopic 25 keV 69Ga+ primary ion source. The primary ion(target) current was typically 2 pA, and the raster area of the beamwas 500 "m × 500 "m. The negative ion spectra were calibratedusing H−, O−, OH−, CH−, CH2

−, and C2H.

2.6. Radio-labeling

Probe oligonucleotides were 3′-end-labeled with [!-32P] dATPusing a Terminal Transferase labeling kit. The reaction mixture con-sisted of 5 pmol of 5′ end amine-terminated oligonucleotide, 5 "l of10× NE buffer 4, 5 "l of 2.5 mM CoCl2, 0.5 "l Terminal Transferase

Page 3: Passivation of aluminum with alkyl phosphonic acids for biochip …test5.coe.utah.edu/files/2014/07/Passivation-of-aluminum.pdf · 2014. 8. 1. · It is known that the phosphonic

7148 S. Attavar et al. / Applied Surface Science 256 (2010) 7146–7150

(20 units/"l), 0.5 "l of 10 mM dATP [!-32P] and ddH2O to a final vol-ume of 50 "l. The mixture was incubated at 37 ◦C for 30 min. 10 "lof 0.2 M EDTA (pH 8.0) was added to terminate the reaction. Theproducts were purified using sephadex g25 columns. The purifiedproduct was spiked with 245 pmol of unlabelled amine-terminatedoligonucelotide. The solution was dried using a speed vac. Driedoligonucleotide was re-suspended in 150 mM phosphate buffer (pH8.5). The silanized substrates were spotted with 1 "l of 50, 5, 0.5and 0.05 "M solution of oligonucleotide, covering three orders ofmagnitude. After spotting, the substrates were held at room tem-perature in a humid chamber for at least 4 h. The substrates werethen rinsed with ddH2O and blown dry with N2. These substrateswere scanned using a phosphor-screen.

3. Results and discussion

Contact angles on cleaned aluminum and glass substrates werenearly zero. Water droplets completely wetted these surfaces.AFM scans of the cleaned Al surfaces showed about 1.2 nm sur-face roughness, which is low enough so as not to affect contactangle [28]. After surface treatment, the aluminum surfaces becamehydrophobic. The contact angle observed for BPA coated aluminumwas 81 ± 2◦, DPA-coated aluminum was 103 ± 2◦ and that for ODPAwas 116 ± 2◦. These values correspond well with the lengths of thealkyl chains of these molecules, with slight aberration in the caseof DPA [29,30]. The contact angle on GPS-silanized glass substrateswas 57 ± 2◦. Nearly the same values were observed with similarlysilanized aluminum substrates: 59 ± 2◦.

Average SAM thickness was calculated based on ellipsometrymeasurements. The phosphonic acid monolayer is modeled as astack of Al/Aluminum oxide/ODPA (or DPA). The aluminum oxidefilm thickness was obtained from ‘! ’ and ‘"’ measured on the ref-erence aluminum oxide/Al stack by using the optical propertiesfrom the standard database (1.75–1.80 in the visible spectrum),in which the only unknown was the thickness of the oxide layer.The thickness of the oxide layer was deduced from fitting ‘! ’ and‘"’ obtained on three spots of this reference film. This value wasused to represent the thickness of the oxide layer in the opti-cal stacks of Al/aluminum oxide/ODPA (and DPA) for the ‘! ’ and‘"’ data collected from three spots on the ODPA and DPA sam-ple.

We adopted the Tauc–Lorentz (TL) generalized oscillator disper-sion model to obtain dielectric properties and thicknesses of thephosphonic acid (PA) monolayers [31]. The TL model combines theTauc joint density of states and a single transition Lorentz oscillatorto account for interband absorption and bound electron absorption,respectively. The dielectric function ε1 + iε2 and photon energy areused in the TL dispersion model.

In the TL model the monolayer thickness was one of the fittingparameters and the dispersion of ‘n’ and ‘k’ were calculated fromthat of the dielectric function by ε1 + iε2 = (n + ik)2. The average alu-minum oxide layer thickness was measured to be 3.77 ± 0.02 nm.The average ODPA monolayer thickness was measured to be1.73 ± 0.24 nm, with a thickness 0.61 ± 0.06 nm for DPA. Similarthickness measurements were reported for monolayer formationof both ODPA and DPA films [32]. Refractive indices (n) in the vis-ible range were 1.82–1.90 for ODPA and 1.62–1.70 for DPA. Therewas no measurable absorbance (k) in all cases.

To further analyze the films, XPS characterization was per-formed to determine the chemical identities of the surfaces. Thesurface chemical composition of clean, unmodified Al was 49.0 at.%Al, 41.5 at.% O, and 8.5 at.% C, with fluorine accounting for 1 at.% orless. Fluorine was present on the aluminum substrates and couldnot be removed using an oxygen plasma cleaning. Based on high-resolution XPS scans of F 1s and C 1s, the fluorine present in the

Fig. 1. High-resolution XPS peak of Al 2p. The peak at 72.9 corresponds with Almetal peak and 75.9 peak corresponds with Al oxide peak.

samples were not in form of fluorocarbons. A high-resolution XPSscan around the Al 2p peak confirmed the presence of an oxidelayer, as shown in Fig. 1.

The Al 2p spectrum was resolved into metallic and oxide com-ponents by fitting in the 65–85 eV binding energy region. The fittedspectrum illustrates the presence of an oxide peak at the bindingenergy of 75.9 eV as well as an Al metal peak at 72.9 eV. This agreeswell with the binding energy separation of 2.8 eV reported in anXPS-spectra handbook. The presence of oxide on the Al surface isrequired for the Al–O–P bond formation.

Fig. 2 shows the survey spectra for ODPA modified aluminum(Fig. 2a) and unmodified aluminum (Fig. 2b). The phosphorus 2sand 2p peaks are indicative of the modification of the Al surfaceby ODPA. Also, C 1s peak suggests the formation of phosphonicacid layer on the aluminum film. These peaks are absent on glassafter similarly treating it with ODPA. High-resolution spectra werecollected at the O 1s, C 1s, P 2p and Al 2p peaks (Table 1).

The C/P ratio for DPA is 10.6. This is close to the theoretical valueof 10. The C/P ratio for ODPA is 21. That this number is greaterthan the theoretical value of 18 is most likely due to attenuationof the P photoelectrons through the moderately thick hydrocarbonover-layer. Adventitious carbon may also contribute somewhat tothis ratio, although its contribution is probably small because ofthe hydrophobic nature of the surface. Similar trends have beenreported in other studies when complete coverage was observed[25].

The stability of these monolayer films was evaluated by dippingthem in both methanol and water for 24 h and then evaluating thesurface coverage. Neither the water contact angles, nor the surfacechemical compositions, changed on the ODPA and DPA modifiedaluminum films. This was not the case for the BPA coated substrates.

Table 1Atomic percentages (±2.5% relative error) were calculated from high-resolution XPSscans.

O 1s C 1s P 2p Al 2p

Al + DPA 28.5 28.5 2.7 39.4Al + ODPA 21.7 51.8 2.5 23.5

Page 4: Passivation of aluminum with alkyl phosphonic acids for biochip …test5.coe.utah.edu/files/2014/07/Passivation-of-aluminum.pdf · 2014. 8. 1. · It is known that the phosphonic

S. Attavar et al. / Applied Surface Science 256 (2010) 7146–7150 7149

Fig. 2. Survey spectra across the binding energy range 0–600 eV. (a) XPS-spectra of ODPA modified Al. (b) XPS-spectra of bare Al.

Their contact angles decreased from 81◦ to 45◦. This behavior is alsocorroborated in the Al/P ratio from the XPS study. The Al/P ratio forODPA is 9.5, for DPA is 14.7 and for BPA is 18.2. The smaller ratio issuggestive of a dense phosphonic acid film. Thus ODPA forms themost dense film among the phosphonic acids used in this study.The samples passivated with BPA had the highest Al/P ratio of 18.2among the phosphonic acids used in this study. This explains theinstability of the BPA film. Studies have shown that the alkyl chainlength has a strong influence on the molecular packing during self-assembly: the longer the chain length, the better the orientation ofthe molecules on the surfaces. The longer chains are better able toself-assemble due to an increase in van der Waals (vdW) attractiveforces with increasing chain length, because the strength of thevdW interactions per adsorbate is proportional to the number ofmethylene units in the adsorbate [2,33].

The resistance of the ODPA and DPA films to silanization withGPS was evaluated by testing the silanized films for Si using XPS.The resulting high-resolution narrow scans over the Si 2s and Si2p regions were at background levels, indicating that there was nosilane detectable on the surface.

Complementary ToF-SIMS measurements were then performedto demonstrate the selective functionalization of the aluminumand glass regions on a prototype biochip. Square windows(200 "m × 200 "m) were opened in 100 nm aluminum films on aglass substrate. These substrates were passivated with DPA or ODPA

and then silanized with GPS. Positive secondary ion spectra failedto show any characteristic peak for the modified aluminum surfacethat was easily distinguishable from the glass region. This was dueto hydrocarbon peaks from the alkyl chains of both the phosphonicacids and the silane, where the lower water contact angle (highersurface free energy) of the silanized glass region would also makeit more susceptible to contamination from adventitious hydrocar-bons. However, characteristic peaks were observed in the negativeion spectra for the phosphonic acids and the silane. Two fragmen-tation peaks of the phosphonic acid group: PO2

− (AMU 63), andPO3

− (AMU 79), confirmed the presence of phosphonic acid on theAl surface. In the case of GPS, a quantifiable peak was observed atAMU 75. There were two possible sources for AMU 75 ion: AlO3

and SiCO2H3−. A negative ion image scan across a raster area of

500 "m × 500 "m showed the localization of AMU 75 in the glassregion, confirming the identity of the ion as SiCO2H3

− (Fig. 3b). ThePO2

− (AMU 63) and PO3− (AMU 79) (Fig. 3a) images were local-

ized to the aluminum regions on the substrate. The phosphate ionfragment peaks signals were almost down to the background levelin the case of glass. This confirms the selective formation of thealkyl phosphonate layer on aluminum acting as a passivation layeragainst the silane.

Radio-labeling experiments showed the effectiveness of thepassivation layer against oligonucleotide adsorption. The four spotson glass in Fig. 4A are the four serial dilution spots, corresponding

Fig. 3. Negative ion images of 500 "m × 500 "m (substrate: GPS/ODPA/Al/glass) (a) image scan for PO2− and PO3

− (b) image scan for SiCO2H3.

Page 5: Passivation of aluminum with alkyl phosphonic acids for biochip …test5.coe.utah.edu/files/2014/07/Passivation-of-aluminum.pdf · 2014. 8. 1. · It is known that the phosphonic

7150 S. Attavar et al. / Applied Surface Science 256 (2010) 7146–7150

Fig. 4. Radio-labeled oligonucleotide spots on silanized slides. The black spots cor-respond to spotted oligonucelotide probe. Lanes: A—glass, B—aluminum, C—ODPApassivated aluminum, D—DPA passivated aluminum. All the substrates have beensilanized with GPS.

(top to bottom) to a serially diluted oligonucleotide at a level of103 ± 10, 85 ± 12, 37 ± 12 and 19 ± 5 fmols. The spots on aluminum(Fig. 4B) correspond to 187 ± 14, 72 ± 25, 57 ± 12, 24 ± 7 fmols.

The slides of phosphonic acid modified aluminum (Fig. 4C and D)show passivation towards silanization and oligonucelotide adsorp-tion; the spots are not visible on the DPA and ODPA modifiedsurfaces. The grey region around the highest concentration spoton the DPA modified surface suggests a slightly higher backgroundas compared with ODPA modified film. The background levels fromradio-labeled oligonucleotide on ODPA passivated correspond to aconcentration of 0.4 fmols. This result shows a passivation ratio ofat least 450:1 when using ODPA passivation. The background lev-els on DPA passivated slides were higher, especially near the higherconcentration spotted region. The background levels on DPA corre-spond to a concentration 0.9 fmols. This result shows a passivationratio of at least 200:1 when using DPA passivation. The BPA mod-ified film was not stable in an aqueous medium; hence it was notconsidered a suitable candidate for passivation studies.

4. Conclusion

In conclusion, phosphonic acid modification is an importantstep in the development of biochips consisting of patterned alu-minum/glass surfaces. In addition, this passivation can work onother metal oxides such as titanium oxide and tantalum oxide. Pas-

sivation is critical for reducing the background signal to low enoughlevels in order to allow effective biochip operation.

References

[1] T. Hauffman, O. Blajiev, J. Snauwaert, C. van Haesendonck, A. Hubin, H. Terryn,Langmuir 24 (2008) 13450–13456.

[2] A. Ulman, Chem. Rev. 96 (1996) 1533–1554.[3] R.G. Nuzzo, D.L. Allara, J. Am. Chem. Soc. 105 (1983) 4481.[4] C.D. Bain, E.B. Troughton, Y.T. Tao, J. Evall, G.M. Whitesides, R.G. Nuzzo, J. Am.

Chem. Soc. 111 (1989) 321.[5] J.C. Love, L.A. Estroff, J.K. Kriebel, R.G. Nuzzo, G.M. Whitesides, Chem. Rev. 105

(2005) 1103.[6] J. Sagiv, J. Am. Chem. Soc. 102 (1980) 92.[7] N. Adden, L.J. Gamble, D.G. Castner, A. Hoffmann, G. Gross, H. Menzel, Langmuir

22 (2006) 8197–8204.[8] S. Tosatti, R. Michel, M. Textor, N.D. Spencer, Langmuir 18 (2002) 3537.[9] E.S. Gawalt, M.J. Avaltroni, M.P. Danahy, B.M. Silverman, E.L. Hanson, K.S. Mid-

wood, J.E. Schwarzbauer, J. Schwartz, Langmuir 19 (2003) 200.[10] R. Helmy, A.Y. Fadeev, Langmuir 18 (2002) 8924.[11] D. Brovelli, G. Hhner, L. Ruiz, R. Hofer, G. Kraus, A. Waldner, J. Schlosser, P.

Oroszlan, M. Ehrat, N.D. Spencer, Langmuir 15 (1999) 4324.[12] G. Hahner, R. Hofer, I. Klingenfuss, Langmuir 17 (2001) 7047.[13] R. Hofer, M. Textor, N.D. Spencer, Langmuir 17 (2001) 4014.[14] M. Textor, L. Ruiz, R. Hofer, A. Rossi, K. Feldman, G. Hhner, N.D. Spencer, Lang-

muir 16 (2000) 3257.[15] J. van den Brand, O. Blajiev, P.C.J. Beentjes, H. Terryn, J.H.W. de Wit, Langmuir

20 (2004) 6308.[16] A. Franquet, H. Terryn, J. Vereecken, Surf. Interface Anal. 36 (2004) 681.[17] O. Cherniavskaya, C.J. Chen, E. Heller, E. Sun, J. Provezano, L. Kam, J. Hone, M.P.

Sheetz, S.J. Wind, AVS (2005) 2972–2978.[18] J. Korlach, P.J. Marks, R.L. Cicero, J.J. Gray, D.L. Murphy, D.B. Roitman, T.T. Pham,

G.A. Otto, M. Foquet, S.W. Turner, Proc. Natl. Acad. Sci. 105 (2008) 1176–1181.[19] G. Saini, R. Gates, M.C. Asplund, S. Blair, S. Attavar, M.R. Linford, Lab Chip 9

(2009) 1789–1796.[20] P.H. Mutin, V. Lafond, A.F. Popa, M. Granier, L. Markey, A. Dereux, Chem. Mater.

16 (2004) 5670–5675.[21] R. Michel, J.W. Lussi, G. Csucs, I. Reviakine, G. Danuser, B. Ketterer, J.A. Hubbell,

M. Textor, N.D. Spencer, Langmuir 18 (2002) 3281–3287.[22] B. Adolphi, J. Anal. Bioanal. Chem. 3179 (2004) 646.[23] I.L. Liakos, R.C. Newman, E. McAlpine, M.R. Alexander, Langmuir 23 (2006)

995–999.[24] M.R. Alexander, G.E. Thompson, G. Beamson, Surf. Interface Anal. 29 (2000)

468–477.[25] E. Hoque, J. DeRose, G. Kulik, P. Hoffman, H. Mathieu, B. Bhusban, J. Phys. Chem.

B 110 (2006) 10855.[26] L. Forget, F. Wilwers, J. Delhalle, Z. Mekhalif, Appl. Surf. Sci. 205 (2003) 44.[27] R. Luschtinetz, A. Oliveira, J. Frenzel, J.O. Joswig, G. Seifert, H. Duarte, Surf. Sci.

602 (2008) 1347.[28] T. Uelzen, J. Mller, Thin Solid Films 434 (2003) 311–315.[29] S. Sun, G.J. Leggett, Nano Lett. 7 (2007) 3753–3758.[30] I.L. Liakos, R.C. Newman, E. McAlpine, M.R. Alexander, Surf. Interface Anal. 36

(2004) 347–354.[31] J.G.E. Jellison, F.A. Modine, Appl. Phys. Lett. 69 (1996) 371–373.[32] A.G. Koutsioubas, N. Spiliopoulos, D.L. Anastassopoulos, A.A. Vradis, G.D. Priftis,

Surf. Interface Anal. 41 (2009) 897–903.[33] D.M. Spori, N.V. Venkataraman, S.G.P. Tosatti, F. Durmaz, N.D. Spencer, S.

Zurcher, Langmuir 23 (2007) 8053–8060.


Recommended