Sunteți pe pagina 1din 21

Progress in Organic Coatings 41 (2001) 99119

Reective and electrically conductive surface silvered polyimide lms and coatings prepared via unusual single-stage self-metallization techniques
Robin E. Southward , Diane M. Stoakley
National Aeronautics and Space Administration, Langley Research Center, Hampton, VA 23581-0001, USA Accepted 12 October 2000

Abstract Highly reective and/or surface conductive exible polyimide lms can be prepared by the incorporation of positive valent silver compounds into solutions of poly(amic acid)s formed from a variety of dianhydrides and diamines. Thermal curing of selected silver(I)-containing poly(amic acid)s leads to cycloimidization of the polyimide precursor with concomitant silver(I) reduction and surface aggregation of the metal yielding a reective and/or conductive silver surface similar to that of the native metal. However, not all silver(I) precursors are effective surface metallization agents and not all poly(amic acid)s metallize with equal facility. Ligand/anion and polyimide structural effects on lm metallization efcacy and on physical properties on metallized lms are reviewed. Published by Elsevier Science B.V.
Keywords: Chemical vapor deposition; Electrically conductive surface; Polyimide lms; Self-metallization; Silver(I) precursors

1. Introduction This review focuses on the fabrication of surface-metallized polyimide lms and coatings by single-stage self-metallizing procedures. The term single-stage refers to metallized lms/coatings which are developed from a single homogeneous solution that contains both the native metal precursor as a positive valent metal complex and the desired poly(amic acid) precursor of the nal polyimide. That is, the polyimide lm is not prepared in a rst stage, and subsequently metallized in a separate and distinct second stage as is the usual route to preparing metallized lms/coatings. The term self-metallization refers to a lm/coating that is cast or applied as a silver(I) ion-doped poly(amic acid) solution after which thermal treatment induces an internal metal ion reduction without the addition of a distinct external reducing agent, e.g., sodium borohydride, hydrazine, UVvis radiation, etc., to give the metallized surface. If the silver(I)-doped lm is cast as the poly(amic acid) (which is usually the case since there are few soluble polyimides), then thermal curing also gives the polyimide via cycloimidization. During the thermal cycle, reduced metal(0) atoms and small clusters formed in the lm aggregate at the surface to give a metallic layer (see Scheme 1 and later discussions). A major goal of the single-stage self-metallization technique is to produce lms/coatings with superb electromagnetic (EM) radiation

Corresponding author.

specular reectivity and surface electrical conductivity via a convenient and efcient route and to produce lms/coatings with unfailing adhesion at the metalpolyimide interface. Silver is the metal of interest to this review since it has a standard reduction potential (E 0 = 0.80 V) which favors facile reduction of silver(I) to metal and since it have has exceptional reectivity and very low electrical resistivity. Polyimides are of interest because of their unusual thermal and oxidative stability and their lm-forming properties. The formation of metallic lms on both inorganic and organic polymeric substrates continues to be of substantial interest. Applications include: (a) contacts in microelectronics [1]; (b) highly reective thin lm reectors and concentrators in space environments for solar thermal propulsion [2], solar dynamic power generation [3,4], and gamma ray imaging telescope systems [5]; (c) the terrestrial concentration of solar energy to generate electric power and process heat [68]; (d) large-scale radiofrequency antennas for space applications [9]; (e) bactericidal coatings [10,11]; (f) surface conductive exible polymeric tapes [12,13]; (g) materials for the construction of adaptive and elastomeric optical devices [14,15]; (h) patterned (chemical, ion beam, plasma, etc.) conductive surfaces on a exible dielectric base [16]. Polymeric supports offer the obvious advantages in weight, exibility, elasticity, and fragility relative to inorganic supports such as glasses, ceramics, or native metals. A natural question which arises at this point is why have several groups made a commitment to investigating and

0300-9440/01/$ see front matter. Published by Elsevier Science B.V. PII: S 0 3 0 0 - 9 4 4 0 ( 0 0 ) 0 0 1 5 6 - 9

100

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

Scheme 1.

developing internal polyimide metallization processes, which are the subject of this review, since there are well-established deposition methods to construct thin metallic lms on a variety of substrates. The answer is processing simplicity and outstanding adhesion at the polymermetal interface. The established methods used to prepare metallized lms require the external deposition of the metallic phase onto the substrate surface. Such methods include: physical vapor deposition via thermal evaporation or sputtering, thermally induced chemical vapor deposition (CVD), photochemically induced CVD, and plasma-assisted CVD, electrodeposition, and electroless chemical reduction of metal from a solution in which the substrate is immersed. Metallization of polymeric lms by these standard deposition techniques involves two, often three, stages. First, the substrate or platform lm is prepared. Secondly, the surface of the base lm is often modied via plasma, ion beam, electron beam, photolytic, or chemical (often oxidation accompanied by etching) treatments to enhance metalsubstrate adhesion. Thirdly, the metal is deposited onto the lm surface via one of the above-mentioned external processes. There are several drawbacks to external deposition procedures with regard to preparing metallized polymeric lms/coatings. The major problem is that the adhesion of more passive metals, such as silver, gold, copper, platinum, palladium, and rhodium, to polymers is notoriously poor [1719]. A second problem with external methods is that they are often not convenient for preparing lms of very large surface area. (Electroless silver deposition Tollens chemistry via spraying can be used to prepare lms of

large area; however, our experience is that the silver layer adheres poorly to polyimides.) Finally, CVD techniques have not been successfully applied to polymers since the surface of the polymer cannot be heated to the temperatures required to reductively decompose volatile metal precursors to the native metal without distortion or decomposition of the base; moreover, the adhesion problem remains. The in situ approach to metallized lms reviewed here leads to the efcient and convenient fabrication of specularly reective and surface conductive metallized lms and should be useful for the preparation of both small and large continuous lms with excellent adhesion at the polymermetal interface. The single-stage self-metallization protocols can be conceptualized as inverse chemical vapor deposition. That is, lm metallization is an internal rather than an external reduction of a metal complex precursor which yields metal atoms and near-atomic metal clusters which diffuse and aggregate to give reective and conductive metallic surfaces with excellent adhesion at the metalpolymer interface. While CVD has limitations for the fabrication of surface-metallized polymeric lms, the chemistry involved in CVD, i.e., thermally promoted metal ion reduction from a molecular precursor, is conceptually apropos to the chemistry involved in the internal polymer metallization described herein. Thus, a limited overview of CVD is appropriate. CVD is an increasingly important route to surface-metallized lms [20,21]. In this technique, metal is externally deposited on a substrate surface by thermal decomposition and positive valent metal reduction from an appropriate volatile molecular precursor. The reduction of

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

101

the metal usually takes place at the substrate surface leaving a metallic thin lm with the ligands of the metal complex precursor being volatilized from the system either intact or in a decomposed form. For silver there is a distinct lack of stable, volatile complexes which can serve as CVD precursors. However, recent progress has been reported by Girolami and coworkers [22,23], Puddephatt and coworkers [2426], and by Kodas and coworkers [27,28], in the synthesis of appropriate silver(I) complexes for CVD and in understanding why earlier silver complexes have not been as volatile as anticipated from copper(I) analogs [28]. Problems are increased for CVD of silver(0) on polymeric substrates since they usually cannot withstand the elevated temperatures (which foster distortion, ow, or decomposition) necessary to reduce and decompose the metal complex. Whereas in CVD a silver(I) complex, such as the monomeric three coordinate (hexauoroacetylacetonato)(triphenylphosphine) silver(I), [(HFA)(PMe3 )Ag] [23] or (hexauoroacetylacetonato)(triethylphosphine)silver(I), [(HFA)(PEt3 )Ag] [24], is volatilized and externally passed over a heated substrate surface, for the metalpolyimide systems reviewed below, positive valent metal complexes, very similar in nature to many CVD complexes of silver and other metals, are added as pure positive valent metal compounds to the polymer solution or are prepared in situ in a polymer solution. A lm is cast or coated onto a glass or metal plate using the metal-doped polymer solutions, and then heated to effect an internal thermal reduction of silver(I). Migration or phase separation of the resultant silver(0) atoms occurs, in part, to give a thin silver(0) lm at the surface of the polymer. A specic example of the single-stage self-metallization approach is illustrated in Scheme 1. A pure isolated silver compound such as silver(I) triuoroacetate, silver(I) nitrate, (1,1,1,5,5,5-hexauoro-2,4-pentanedionato) (4 -1,5-cyclooctadiene)silver(I), etc. is dissolved in a solution of a poly(amic acid); alternatively, a silver(I) complex can be prepared in situ without isolation in the solvent in which the poly(amic acid) resides. An example of the in situ approach is also shown in Scheme 1. Specically, silver(I) acetate is allowed to react with the -diketone 1,1,1,5,5,5-hexauoro-2,4-pentanedione (hexauoroacetylacetone or HFAH) and 1,5-cyclooctadiene in an appropriate solvent such as dimethylacetamide (DMAc). This reaction generates the (1,1,1,5,5,5-hexauoro-2,4-pentanedionato) (4 -1,5-cyclooctadiene)silver(I) complex, Ag(HFA)(COD), which is not isolated from solution, by enol proton transfer from the HFAH ligand to the acetate anion of silver(I) acetate. A solution of the poly(amic acid) in DMAc is added to the solution of Ag(HFA)(COD). A lm is then cast. Heating the silver(I)-doped poly(amic acid) lm effects reduction of silver(I) to the native metal as solvent and water of imidization are lost from the system in forming the metallized polyimide lm. During the thermal cycle a portion of the silver(0) formed initially as atoms, then very small clusters, in the matrix of the polymer, aggregates at the lm surface

to give a thin metallic layer. Depending on the system highly reective and/or conductive surfaces are formed. One of the expectations was that adhesion would be strong due to mechanical interlocking, and indeed, this has been realized. Another expectation was that the metallized lms would retain the essential mechanical properties of the metal-free parent polymer. This seemed a reasonable hope with silver(0) since, as a more passive metal, it does not interact strongly with polymer functionalities. Again, as detailed later, for selected systems metallized polymeric lms are produced with excellent mechanical properties. Aromatic poly(amic acid)-polyimide systems have dominated the area of self-metallizing lms because of their outstanding chemical and thermal stability and excellent mechanical properties [29,30]. Thermal stability is desirable since silver catalyzes air oxidation of organic functionalities at the temperatures required to produce metallized lms [31]. The metallized lms of this review have been characterized by a multiplicity of techniques including: differential scanning calorimetry (DSC), thermal gravimetric analysis (TGA), thermal mechanical analysis, reectivity measurements, resistivity measurements, X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), transmission electron microscopy (TEM), atomic force microscopy (AFM), X-ray diffraction, mechanical testing, and chemical analysis. 2. In situ surface metallization of polyimides with silver 2.1. General considerations of silver The specular reectance of metallic silver over the solar spectrum exceeds that of all other metals [8]. Silver is a metal of choice as a reecting material because of its singularly high reection coefcient (0.93) and low coefcient of solar absorption (0.07). The specular reectance of silvered mirrors is unexcelled, and the hemispherical reectance of a clean silver lm is greater than 97% weighted over 2502500 nm, the range of the solar spectrum. Silver also has the highest electrical conductivity of all metals at 6.3 107 ( m)1 , and it is relatively inexpensive. For these reasons, techniques for fabricating silvered lms have been investigated for numerous applications as cited earlier. While the reectivity and electrical conductivity of aluminum is excellent, and while aluminum is chemically resistant due to a very ne oxide coating, the highly negative standard electrode potential of aluminum makes it unsuitable for the electroless reductive self-metallization processes reviewed herein. However, there are disadvantages to using silver as the reecting metal. First, silver is a relatively soft metal so that the face of a mirror needs to be carefully protected from mechanical abrasion. Second, silver tends to tarnish which diminishes its reectivity. (Ambient sulfur-containing compounds are a particular problem.) Third, and perhaps most importantly, silver(0), as a more passive metal, does not interact strongly with organic

102

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

functionalities, which means that adhesion of a silver layer on a polymer surface can be a substantial problem as mentioned before [1719,3235]. Thus, in selected applications it may be necessary to protect the surface of a silver mirror with an appropriate top-coating. Generating sufcient adhesion between the metal and the base polymer has proven to be a more challenging problem that has not been solved for traditional deposition techniques as mentioned earlier. 2.2. First metallization studies To our knowledge, the rst work on in situ reduction of silver(I) in polyimide lms was reported in the patent literature by Endrey [13] in 1963 using silver(I) carboxylate complexes. This effort is disappointing since much is

claimed in the midst of a dearth of characterization data. A large number of silver(I) salts of oxy acids of carbon, e.g., formic, acetic, propionic, succinic, oxalic, maleic, furmaric, citric, etc., were said to have been added to organic solutions of a variety of polyimides. However, in the four examples given in the patent, the silver salt was either silver acetate or silver caprylate, the organic solvent was dimethylformamide (DMF), and in every case excess pyridine was added to aid in dissolution of the silver carboxylate complex and to keep the poly(amic acid) from gelling. Only two poly(amic acids) were reported in the examples: the ones formed from pyromellitic dianhydride (PMDA) with 4,4 -diaminodiphenylmethane and with m-phenylenediamine (see Scheme 2 for monomer structures). Endrey states that addition of the silver salts to the poly(amic acid) leads to a

Scheme 2.

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

103

metathesis reaction which gives a poly-silver(I) salt of the amic acid. This is reasonable since an amic acid carboxylate group is less basic than an acetate or caprylate group. Such complexation tends to give a gelled system since it is established that silver carboxylate complexes are dimeric involving the coordination of two carboxylate groups with each silver(I) ion [36,37]. Gelation was broken down by overwhelming the system with a silver(I) complexing agent, specically pyridine. Pyridine is well known to form a linear bis complex with silver(I) [38]. Endrey states that more solvent or pyridine or a -ketonic type compound such as ethyl acetoacetate are advised to clear any gel or insoluble matter that may form in the polyamide-acid salt solution. Specically, a clear viscous dope was formed with rapid stirring upon the addition of a solution of silver(I) acetate, dissolved in a large excess of pyridine, to a DMF solution of the poly(amic acid). The polyamide-acid salt was then cast onto a glass plate and thermally cured in a vacuum oven to 300 C where the lm was held for 30 min. A polyimide lm was produced with embedded metallic silver particles as shown by X-ray diffraction. Some lms were visually clear which was interpreted to mean that the particles therein do not have dimensions greater than 0.8 mm. Initially, the metallized lms were not conductive. However, upon further heating in nitrogen for several hours at 275 C an electrically conducting surface was achieved. No measurement or mention of the reectivity of the lms was made, and no mechanical or thermal data were mentioned. No journal publications by Endrey or others at DuPont elaborating the work in the patent have appeared. No examples of lms prepared with -diketonic-type compounds were presented in the examples, even though compounds were claimed to be solubilizing, gel-preventing ligands. Thus, there is substantial uncertainty as to the full range of properties of the Endrey lms. 2.3. Silver(I) nitrate studies A surprising aspect of the Endrey patent is that silver(I) nitrate is not mentioned as a source of homogeneous silver(I) polymer solutions and lms. Silver(I) nitrate is available in high purity at a modest cost; it is soluble in DMAc, N-methylpyrrolidinone (NMP), and DMF, which are common and useful solvents for the management of poly(amic acid) precursors. The rst report of the use of silver(I) nitrate in a poly(amic acid) was by St. Clair and Taylor [39] in 1983 with additional, but limited, silver(I) nitrate studies appearing through 1992 [3943]. These investigators examined silver(I) nitrate as the source of silver(0) in lms of the traditional poly(amic acids) BDTA/4,4 -ODA, BDSDA/4,4 -ODA, and PMDA/4,4 -ODA (see Scheme 2 for structures). Thermal curing of the doped amic acid precursors gave lms which varied in reectivity with some being verbally described as very reective. BTDA/4,4 -ODA lms were prepared with silver(I) nitrate at 1.3, 4.0, 5.3, and 10.0% silver. In general, silver(I) nitrate lms were found

not to be electrically conductive at the surface; they are not conductive in the bulk since concentrations are below the percolation threshold. Unfortunately, quantitative data were minimal. The lack of surface conductivity was ascribed to a polymer overlayer which could be removed partially by sputtering with argon. It was claimed that if silver(I) nitrate was added in the absence of light, a conductive lm resulted [44]. No conrmation of this observation has appeared in the literature, and we have had no experience which would suggest an explanation for this result. The only quantitative reectance work with silver(I) nitrate found that the reectivity was maximized in BTDA/4,4 -ODA when the silver(0) concentration was ca. 5% and the lm was post-cured or sintered at 300 C for 8 h [43]. (Above ca. 5% silver the lms are seriously degraded on thermal curing.) In this single case the reectivity was maximized at 48%. It is clear from thermal and mechanical data that silver(I) nitrate adversely affects polyimide properties at concentrations greater than ca. 45% silver. Tg for a 4.0% silverBTDA/ODA lm was 320 C while that for the parent polyimide was 286 C; for a 5.3% lm Tg was 332 C. This suggests that the presence of silver(0) and/or silver(I) in the cured polyimide induces crosslinking. The mechanical properties were degraded with the tensile strength of the 4.0% doped lm at room temperature being 12.0 ksi and the modulus being 215 ksi compared with 16.5 and 284 ksi, respectively, for the undoped polyimide. The presence of silver(0) in polyimide lm reduces thermal degradative stability in air. The 10% weight loss value using dynamic thermal gravimetric analysis (TGA) was ca. 400 C compared with ca. 550 C for the parent lm. Thus, it appears that silver metal catalyzes oxidative degradation as would be expected since there is a rich literature on silver catalyzed organic reactions [31]. Often the lms were brittle and could not be removed from the casting plates without disintegrating. Generally, lms had lower tensile strengths and higher moduli when these could be measured. XPS data showed that there was much more silver on the air side of the lm for BTDA/4,4 -ODA than with silver(I) nitrate in BDSDA/4,4 -ODA even though Tg for the latter polyimide is 70 C less than for BTDA/4,4 -ODA. It was suggested that the coordination of silver to the thioether groups may account for this reduced migration. Comparison of binding energies and concentrations at the metallized air side surface for C, N, and O of the parent and metallized BTDA/4,4 -ODA lms showed little difference. This nding was interpreted to mean that even though there is considerably more silver on the air side, it must be either segregated on the polyimide surface or present as a very thin lm . . . . The persistence of polyimide near the surface prevents the silver particles from coming into contact to form a conductive layer. Independent of the above work, Auerbach [45] in 1984 reported in a brief communication the use of silver(I) nitrate in a poly(amic acid), the classic and prototypical poly(amic acid) synthesized from PMDA and 4,4 -ODA, which led to conductive lms with NMP as the solvent.

104

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

During the thermal cure cycle the lm was bathed in carbon powder which was said to accelerate the reduction of the silver(I) to metallic silver. The formation of the silvered surface appeared to be sensitive to the presence of adequate oxygen since it was observed that, if too much solvent remained in the poly(amic acid) cast lm, reduction to silver metal was inhibited. This effect was ascribed to the evaporation of the solvent from the polymer during curing which prevented oxygen from reaching the polymer surface during imidization. More likely (vide infra), oxygen is needed to degrade surface polyimide which allows silver particles to come into contact. Auerbach also found that too much silver(I) nitrate compromised the metallized lm quality, although no stoichiometric information was presented to know what too much means. The optimum lms were reported to be conductive and reective, but strangely no quantitative data were given. Finally, Auerbach [46] reported that silver(I) reduction in a silver(I) nitrate doped poly(amic acid) lm could be achieved with a low power laser, which was said to have implications for writing conductive pathways onto lm surfaces. One presumes that this was a thermal rather than a photochemical stimulated reduction; again quantitative data were inadequate. Southward et al. [47] in 1995 reported that a 5.4% silver BTDA/4,4 -ODA lms prepared with silver(I) nitrate and cured to 300 C in air had a reectivity of 20% and was not conductive. They found that lms with silver concentrations beyond 5.4% were visually degraded and brittle without mechanical usefulness. 2.4. Silver(I) carboxylate, organosulfonate, sulfate, tetrauoroborate, oxide, and trimethylphosphineiodosilver(I), [(PMe3 )AgI]4 , studies Boggess and Taylor [42], in addition to silver(I) nitrate, also utilized silver(I) benzoate, silver(I) triuoromethanesulfonate, and silver(I) sulfate as additives to selected poly(amic acids). The silver(I) benzoate and sulfate in BDSDA/ 4,4 -ODA (see Scheme 1) gave lms that were described as transparent and red; silver(I) triuoromethanesulfonate was brown and nearly opaque. Elemental analysis for silver in the bulk of these lms gave values which were often very low compared with theoretical values, and there was considerable variation from sample to sample. No explanation was given for these analytical results, but they suggest that there may be silver(0) catalyzed oxidative degradation of polyimide. Tg for the BDSDA/4,4 -ODA-silvered lms was close to that of the parent polymer at concentrations near 4% silver. None of these lms was described as even modestly reective or conductive. Southward et al. [47] found

that silver(I) tetrauoroborate at 10% silver gave a lm that was 48% reective with better mechanical properties than a silver(I) nitrate lm. This suggests that the nitrate anion may be primarily responsible for lm damage in the silver(I) nitrate system. When insoluble silver(I) oxide is dispersed in a BTDA/4,4 -ODA poly(amic acid) matrix and cured, the oxide goes completely to metallic silver as shown by X-ray diffraction data, but the lm has no metallic surface (%R = 5%) [47]. The large aggregates of silver(0) formed from the micron size silver(I) oxide particles are too large to undergo migration to the surface in the polymer matrix. [(PMe3 )AgI]4 , which has the cubic structure, is soluble in DMAc. When incorporated into a BTDA/4,4 -ODA lm, the cured lm is not reective (7%) and both silver(I) iodide and metallic silver are present [47]. Silver(I) iodide comes about from thermal release of trimethylphosphine near 80 C. Thus, from this earlier work (19831995) it appeared that the simplest or most elementary silver(I) salts such as carboxylates, nitrate, sulfate, oxide, etc. (the sulfate and oxide being insoluble in the resin solution) give lms which exhibit, at best, modest specular reectivity and are not conductive. However, recent results show that this is not the case for all carboxylate complexes as seen in the following paragraph. In 1998, Southward et al. [48,49] reported that silver(I) peruorocarboxylate salts, AgCn F2n+1 CO2 , n = 13, particularly silver(I) triuoroacetate with n = 1, with BTDA/4,4 -ODA gives metallized lms with excellent specular reectivity, some of which become conductive with light polishing, i.e., gentle abrasion. Silver(I) triuoroacetate lms were studied over a concentration range 411% silver (see Table 1 for selected examples). Silver lms at 10.7% metal were cast on a glass plate or on an undoped parent polyimide base and cured to 300 C for 1 h in air. They were 65% reective, but not conductive as taken from the oven and measured by a four-point probe. However, these lms could be gently rubbed with a soft polishing cloth to give a reectivity greater than 97% and a conductivities of 19 and 8 /sq, respectively. A sample with only 6% silver, cured to 300 C for 1 h, achieved an unpolished reectivity as of 83%. Curing to 340 C did not improve lm properties. XPS data are consistent with residual surface polyimide which is in a partially degraded state. The C to O ratio is that of the parent BTDA/4,4 -ODA, but N is low. The authors suggest that residual partially degraded polymer at the surface limits the reectivity via absorption and scattering and keeps silver(0) particles from coming into contact to give a conductive surface. Gentle abrasion removes surface polymer residue and enhances contact among silver(0) particles.

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

105

106

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

Reduction of silver(I) triuoroacetate in BTDA/4,4 -ODA does not lead to any change in Tg , and the mechanical properties of the AgCF3 CO2 silvered lms are similar to those of the parent polyimide. These ndings are in stark contrast to the observations with silver(I) nitrate in BTDA/4,4 -ODA where Tg is elevated by ca. 50 C and mechanical properties are degraded. SEM reveals that the 6.0% lm of Table 1 has surface silver particles sizes of 125200 nm and that the particles, while partially interconnected, are overall isolated from one another, thus explaining the lack of conductivity. When a 10.7% lm (not listed in Table 1) is cured at 135 C for 1 h, heated over 4 h to 300 C, and just reaches 300 C, the particle morphology is well established, and a reectivity of 81% is achieved. The predominate particle sizes are in the range 125200 nm, similar to the 6.0% lm but less densely distributed. After remaining at 300 C for 7 h, there is still a preponderance of isolated particles, and the lm is not conductive; however, the average particle size increases to 150300 nm. The shapes of the particles after 7 h at 300 C suggest that there is an increase in size due to sintering. Several apparent necks are visible in the micrographs. Thus, particle size appears to be function of concentration and the details of the cure cycle. Additional studies with silver(I) triuoroacetate with poly(amic acid)s are clearly warranted and may yield exceptional lms. In particular, as will be seen in systems discussed later, higher concentrations of silver (ca. 13%) and longer cure times may lead to lms which are both highly reective and conductive.

2.5. (1,1,1,5,5,5-Hexauoroacetylacetonato)(4 -1,5cyclooctadiene)silver(I) studies From 19941998, Taylor and coworkers [5054] published noteworthy single-stage self-metallization work using (1,1,1,5,5,5-hexauoroacetylacetonato)(1,5-cyclooctadiene) silver(I), Ag(HFA)(COD), as the metallic silver metal precursor (this complex is commercially available). Pertinent results are summarized in Tables 2 and 3. In the solid state, Ag(HFA)(COD) is a dimeric complex, with unusual bridging -diketonate ligands [55] as shown below.

The complex is soluble in DMAc and in poly(amic acid) matrices. In DMAc, Ag(HFA)(COD) may well be monomeric since oxygen bridging acetylacetonate ligands are not common, and high concentrations of a polar donor solvent such as DMAc may solvate and stabilize monomeric structures. The rst [5052] poly(amic acid) systems which

Table 2 Selected reectivity, resistivity, and thermal data for polyimides doped with [(HFA)(1,5-COD)Ag] (cured at 100 C for 1 h, 200 C for 1 h, 300 C for 1 h) Filma Polymer:dopant mole ratio (ca. %Ag)b Percent of reectivityf (at 531 nm) 20 BTDA/4,4 -ODA BTDA/4,4 -ODA BTDA/4,4 -ODA BTDA/4,4 -ASD BTDA/4,4 -ASD BTDA/4,4 -ASD DSO2DA/4,4 -ODA BTDA/DDSO2 BTDA/4,4 -ODA/4,4 -ASD BTDA/APB ODPA/4,4 -ODA BDSDA/4,4 -ODA BDSDA/4,4 -ASD BPADA/4,4 -ODA BPADA-4,4 -ASD
a b

Resistivity ( /sq)d

Tg ( C) Parent Doped 265 265 263 268 270 262 293 249 272 280 268 221 221 227 225

10% weight losse Parent 609 609 609 608 608 608 584 619 608 638
c

45 57 50 44 51 40 36
c c

70 45 32 22 49 31 18
c c

Doped 493 493 481 516 482 478 474 586 502 478
c

2:1 4:1 8:1 2:1 4:1 8:1 2:1 4:1 2:1 2:1 2:1 2:1 2:1 2:1 2:1

(10% Ag) (5.3% Ag) (2.7% Ag) (9.7% Ag) (5.1% Ag) (2.6% Ag)

65 60 53 55 52 46
c c

>1011 NC NC >1011 NC NC NC NC NC NC NC 320 104 >1011 104 (50)

(9.8% Ag) (8.5% Ag) (7.4 % Ag) (7.2 % Ag) (xx% Ag) (xx% Ag)

60 56 49 4 4 55 5

52 45
c

48 24
c

3 3
c c

9 7
c c

270 270 270 270 270 270 296 273 270 276 264 214 218 219 217

598 566
c

502 469
c

584

463

All lms were cast and cured on glass plates. Ag is calculated on the basis of only silver(0) and polyimide remaining in the cured lms. Ligands and ligand fragments are assumed to be volatilized from the lm. c Not reported. d Four-point probe. e Heating rate not reported. f Reectivity data arc relative to a Perkin-Elmer polished aluminum optical mirror set at 100%.

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119 Table 3 Mechanical data for selected polyimide lms of Table 2 doped with [(HFA)(1,5-COD)Ag] (cured at 100 C for 1 h, 200 C for 1 h, 300 C for 1 h) Filma Polymer:dopant mole ratio (ca. %Ag)b Tensile strength (ksi) Parent BTDA/4,4 -ODA BTDA/4,4 -ASD BTDA/4,4 -ODA/44 -ASD BDSDA/4,4 -ODA
a b

107

Modulus (ksi) Parent 445 379 409 209 Doped 317 343 413 153

Doped 11.8 15.1 16.1 6.10

2:1 2:1 2:1 2:1

(10% Ag) (9.7% Ag) (9.8% Ag) (7.4% Ag)

17.9 16.5 17.3 12.7

All lms were cast and cured on glass plates. Ag is calculated on the basis of only silver(0) and polyimide remaining in the cured lms. Ligands and ligand fragments are assumed to be volatilized from the lm.

were doped with Ag(HFA)(COD) to develop reective lms included: BTDA/4,4 -ODA, BTDA/4,4 -ASD, BDSDA/ 4,4 -ODA, BDSDA/4,4 -ASD, BTDA/APB, BTDA/DDSO2 , DSO2 DA/4,4 -ODA, and the random copolymer BTDA/4,4 ODA/4,4 -ASD (see Scheme 2 for structures). Thermal curing to 300 C (usually in air) of these poly(amic acid) lms doped with Ag(HFA)(COD) (ca. 319% silver) gave metallized lms which had air side reectivities of 465%. Most lms were not conductive. For all samples cured on glass plates, XPS revealed that the air side of the lm had 1050 times more silver than the glass side; the glass sides of the lms were never reective. The specular reectance was highest (65%) with BTDA/ 4,4 -ODA lms, with the reectivities for BTDA/4,4 -ASD, BTDA/3,3 -APB, and the BTDA/4,4 -ODA/4,4 -ASD copolymer lms being in the range 5560%. For all lms the reectivity decreased noticeably with increasing angle of incidence. For the 2:1 (polyimide repeat unit to silver ratio) BTDA/4,4-ODA and BTDA/4,4 -ASD lms, the theoretical weight percent silver based on polyimide and silver alone (i.e., assuming that the HFA and COD ligands were lost from the system in some form) remaining after thermal curing should have been 10.0 and 9.7%, respectively. The amount of silver actually found after thermal curing was 8.4 and 8.1%, respectively. This is in rough agreement with that calculated on the basis of only metallic silver and polyimide remaining after curing. The values found are lower than those calculated in part due to residual HFA in some form residing in the nal polyimide as indicated by the fact that uorine was found to be 2.2 and 3.6%, respectively. The analytical results indicate clearly that the metallic silver formed in the lms does not catalyze oxidative degradation of the polyimide during thermal curing. Such degradation would not be surprising since silver is a well-established oxidation catalyst for organic molecules. At an 8:1 ratio, the two cured BTDA/4,4 -ODA and BTDA/4,4 -ASD lms have calculated silver concentrations of 2.7 and 2.6%, respectively. Surprisingly, the reectivities did not diminish in proportion to the reduction of silver and remained relatively high at 53 and 46%. None of the BTDA/4,4 -ODA lms was reported to be electrically conductive; however, the BDSDA/4,4 -ODA and

BTDA/4,4 -ASD lms were surface conductive with sheet resistivities of 320 and 1.34.2 104 /sq as taken from the oven, and after polishing 3 and 20 /sq, respectively. The authors note that the lms with the greatest conductivity were the ones with the poorest reectivities, which they ascribed to increased surface roughness of the conductive silver(0) layer. XPS data for the 2:1 BDSDA/4,4 -ODA lm showed extensive formation of sulfuroxygen linkages in the sulfoxide/sulfone/sulfate region. Thus, it is clear that sulfur is being oxidized, at least at the surface. The formation of SO linkages in the polymer backbone could also contribute to the somewhat lower silver values found from elemental analysis data. Milling experiments would be interesting to see if there was signicant sulfur oxidation in the bulk of the lm. The authors do not make clear whether it is silver(I) or silver metal that is causing the oxidation of sulfur. They state silver also appears to interact with thioether groups in the polymer backbone thereby causing oxidation to a sulfone-like moiety [54]. It is conceivable that Ag(I) is actually oxidizing the sulde to a sulfoxide or sulfone, and that this is the mechanism for the production of metallic silver. Further studies with sulde-containing polyimides and other silver(I) additives should be pursued. As might be expected, the presence of metallic silver adversely affects the thermal stability of the polymers in air as seen from the data for 10% weight loss in Table 3. Thermal stability (10% weight loss) in air for silvered lms is lower by ca. 100 C than that for the parent polyimide. Nonetheless, the metallized polyimide lms still have a wide thermal use range. The authors found that within experimental error Tg did not appear to change . . . [54]. Mechanical properties for the Ag(HFA)(COD) lms were often similar to those of the parent polyimide for several systems in as seen in Table 3. Indeed, all of the lms could be folded on themselves without fracture. Thus, Ag(HFA)(COD) gives lms with much better thermal properties (Tg and 10% weight loss) and mechanical properties (tensile strength and modulus) than silver(I) nitrate. These results, coupled with the silver(I) triuoroacetate lms discussed above, demonstrate that ligand effects are pronounced in the fabrication of metallized lms. Taylor and coworkers [51] found that curing in nitrogen did not alter signicantly the properties of metallized lms prepared with Ag(HFA)(COD) and BTDA/4,4 -ODA.

108

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

The SEMs for the sulfur-containing BDSDA/4,4 -ODA (2:1) and 4,4 -ASD (2:1) lms showed much larger silver particles (100500 nm) and less uniformity among particle distribution than that observed for the corresponding BTDA lms with particle sizes ca. 100 nm and uniformly dispersed at the surface. From TEM the thickness of the 2:1 BTDA/4,4 -ODA silver layer was 70 nm whereas for the corresponding sulfur-containing BDSDA/4,4 -ODA lm it was 140 nm. These observations seem to be reversed with silver(I) nitrate as the dopant as cited earlier. In the nitrate case, Taylor and coworkers found from XPS data that the BTDA/4,4 -ODA lm had much more silver on the air side than the BDSDA/4,4 -ODA analog. This variance was not addressed by the authors. The BTDA/4,4 -ODA silver lm (2:1), which is not conductive, clearly has silver particles which are surrounded by polymer which insulate the metal clusters from one another. This was demonstrated by microscopy data which showed that the silver(0) layer was not continuous and from the XPS which showed the atom percents of carbon, nitrogen, and oxygen to be near 59, 6.4, and 14%, respectively. These atom percentages are approximately what is expected in a parent BTDA/4,4 -ODA lm. Thus, silver(0) is not catalyzing surface polyimide oxidative degradation, and silver(0) migration is not sufcient to cover the surface so that particles are in contact to make the lm conductive. Several conclusions can be drawn from this 19941996 Ag(HFA)(COD) work. First, it is possible to develop a uniform surface layer of silver metal, which exhibits modest reectivity, on a polymer lm by thermal reduction/decomposition of a dissolved silver(I) coordination compound. Secondly, lms that were found to be highly reective were not conductive, whereas the lms which exhibited no metallic reectivity were conductive. There appears to be an inverse relationship between conductivity and reectivity. Thirdly, the presence of sulfur linkages in the polyimide repeat unit leads to larger surface particles of silver; however, the larger the surface particles, the lower is the specular reectivity. The authors conclude that the development of the surface layer is a complex and subtle function of numerous variables including polymer structure, thermal cure cycle, the amounts of oxygen and moisture in the cure cycle atmosphere, the solvent, and the coordination environment of the silver. Even though there are a multiplicity of variables affecting metallization, they felt able to hypothesize that electrically conductive lms (which were not specularly reective), would result if the polyimide exhibited low Tg (ca. 200 C) and contained sulfur and that specularly reective lms (which were not conductive) would result if sulfur were absent in the polyimide regardless of Tg . The conclusions of the authors were further supported in two subsequent papers [53,54] where three additional metallized polyimides (BPADA/4,4 -ODA, BPADA-4,4 -ASD, and ODPA/4,4 -ODA see Scheme 2) were studied with Ag(HFA)(COD) and previously studied systems were examined further. Table 2 shows that for the reectivity and

resistivity data for these additional polyimides. Consistent with their hypothesis the new sulfur-containing polyimide BPADA-4,4 -ASD with a low Tg (217 C) exhibited low resistivity and was not reective; on the other hand, the non-sulfur-containing analog, BPADA/4,4 -ODA, had a reectivity of 55%, but was not conductive. This was consistent with the previous ndings that conductive lms have low reectivities. Comparison of BTDA/APB and BDSDA/APB lms via TEM showed again that the conductive sulfur-based lm had larger and more irregular silver particles at the surface. This was also true for the BDADA/4,4 -ODA and BDADA/4,4 -ASD pair where the conductive silvered BDADA-4,4 -ASD lm showed larger and less uniform metal particle size. Ion milling Auger spectra for silvered conductive BDSDA/ASD and silvered non-conductive BTDA/4,4 -ODA lms show that the silver concentration increased 100% after milling for a short time. This indicates that there is a greater excess of polyimide or polyimide fragments very close to the surface. (This was also seen by Southward et al. with (triuoroacetylacetonato)silver(I) in BTDA/4,4 -ODA; vide infra.) The suggestions of the Taylor group concerning the importance of sulfur in polymers and the glass transition temperatures are interesting, but the study of additional systems and more characterization data are desirable. There are several matters which require attention. First, it is important to follow the production of silver metal during the thermal cure. This can be done by looking at the development of specular reectivity and X-ray diffraction peaks for f.c.c. silver as a function of time and temperature during the cure cycle. X-ray diffraction gives information both with regard to the formation of a metallic silver phase and to the size of the crystallites. The mechanism of silver(I) reduction remains unclear. The electron for silver(I) reduction might come from the ligand, from solvent, from polymer, or from oxygen in Ag2 O formed in situ during the cure cycle. If the electron comes from the polymer, which seems entirely possible, then changing polymer structure could have a dominant effect, that is more important than Tg , on the resulting lms in terms of reectivity, particle size, and conductivity, Tg values for the nal polyimides may well have little effect on the overall migration processes since migration may occur for the most part while the polymer is predominantly in the amic acid form adulterated with residual solvent. Tg values for poly(amic acids) are much lower than for the corresponding polyimides, and even lower where plasticized with solvent and the silver(I) additives. Thus, silver(0) atom/cluster migration may be completed by the time that mobility would be governed by Tg of the nal polyimide. Secondly, it is essential to have both SEM and TEM micrographs as a function of the cure cycle because a lm which is very low in reectivity may have a similar amount of surface silver as one high in reectivity with the reduced reectivity being due to degradation contaminants or roughness at the surface. Microscopy would also give denitive information concerning the details of silver(0) aggregation. It is

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

109

conceivable that conductivity is governed by both silver(0) aggregation and surface degradation of residual polyimide which allows the silver particles to come into contact. Indeed, we have found this to be the case in the (triuoroacetylacetonato)silver(I) BTDA/4,4 -ODA system discussed below. Thirdly, it is a very appealing idea to have low valent sulfur in the poly(amic acid)-polyimide systems since sulfur could be the source of the electron for silver(I) reduction, the sulde then being oxidized to a sulfoxide and sulfone. Thus, oxidized polyimide as a sulfoxide or sulfone would be entirely stable, i.e., we would have the polymer intimately involved in electron transfer to form silver metal, yet the polyimide would not suffer any essential damage as the polymer oxidation product itself would be a stable polyimide. 2.6. Studies with (1,1,1,5,5,5-hexauoroacetylacetonato)silver(I) prepared in situ One drawback to the use of Ag(HFA)(COD) is its instability. The complex loses olen slowly when exposed to air over long periods of time and should be stored in a tightly closed container and kept in a refrigerator according to Partenheimer and Johnson [56], who reported its synthesis. Also, Ag(HFA)(COD), alone and in curing lms, has a obnoxious stench associated with the readily liberated 1,5-cyclooctadiene ligand. Furthermore, when we duplicated the work of Rubira et al. [50,51], we found that the metallizing lms (ca. 225 cm2 in area) detached from the glass plate, except at the very edges, at ca. 200 C during the thermal cure and inated to form a single dome-like structure. Ination is presumably due to the release of volatile solvent and/or gases from ligand decomposition. These gases create pressure against the exible mixed poly(amic acid)-polyimide lm. Ination of the silverpolymer lm is facilitated by partial metallization of the glass side of the lm which decreases adhesion of the curing doped lms to the glass plate. Thus, large area lms prepared with Ag(HFA)(COD) give unsightly and deleterious crease lines as the weight of the metallizing lm cannot be supported by the pressure of the escaping gases. However, reectivities of ca. 65% with the Ag(HFA)(COD)BTDA/4,4 -ODA system were encouraging, and AgHFA-based complexes seemed to warrant further study. Since silver(I) complexes are often unstable both thermally and photolytically, Southward et al. [47,5759] suggested that better metallized lms might be realized if the silver(I) complex could be freshly prepared for each lm preparation. Further, they reasoned that this could be done most easily by preparing the complex in situ in the same solvent in which the poly(amic acid) was dissolved (usually DMAc). An example of such an in situ complex preparation was described earlier and summarized in Scheme 1 for Ag(HFA)(COD). The initial work of Southward et al. [47,57] involved preparing AgHFA without COD, trialkarylphosphines, or any other stabilizing ligands apart from what

interaction the solvent, DMAc, or polymer donor groups might play in coordinating to the silver(I)-HFA species. Southward et al. [47,5759] rst studied BTDA/4,4 -ODA metallization with the in situ preparation of (hexauoroacetylacetonato)silver(I), without the COD ligand, in DMAc (see Scheme 1). It was found that silver(I) acetate, which is insoluble in DMAc, dissolved within minutes after adding ca. 1.1 equivalents of hexauoroacetylacetone (HFAH). Addition of BTDA/4,4 -ODA (1215% by weight in DMAc) to the solution of AgHFA gave a clear, homogeneous, yellow silver(I)-doped solution. Doped poly(amic acid) lms were cast at thicknesses (ca. 375500 mm) to give cured metallized lms near 25 mm. Section A of Table 4 shows that lms produced with 5.312.8% silver are very reective, with the maximum reectivity being significantly greater than that reported by Taylor and coworkers with the isolated Ag(HFA)(COD) complex. The reectivity decreases with increasing angle of incidence, which may be mostly determined by the presence of signicant amounts of polymer at the surface which strongly absorbs in the visible. Tg for the metallized lms is close to that for the parent polymer as are the moduli and tensile strengths. This indicates that there is minimal damage to the essential polyimide structure when silver(I) undergoes reduction upon heating to 300 C. None of the lms was conductive even when the silver concentration was increased to ca. 18%. Curing under nitrogen gave essentially similar lms as seen in Section B of Table 4. Also, the use of silver(I) uoride rather than silver(I) acetate gave similar lms as seen in Section D of Table 4. SEMs showed that there were uniform globular particles of silver metal formed at the air side surface of the lms with particle sizes of ca. 6070 nm. This is very similar to the result reported by Taylor and coworkers using solid Ag(HFA)(COD), but better reectivity was realized. The particle morphology is similar to what occurs via the island growth mechanism with vapor deposition techniques. The lack of conductivity is consistent with polyimide separating the metal particles. SEM data shows clearly that silver particles are not in contact. Holding the lm at the nal 300 C cure temperature for 7 h does not induce conductivity. Southward et al. with Taylor and coworkers suggest that there is an overlayer of polyimide at the surface. XPS data always show substantial C, N, and O at the surface in a atom ratio similar to the parent polymer. This overlayer would account in part for the angle dependence of the reectivity as the pathlength through absorbing surface polyimide increases with the angle of incidence. Also, the absorption of light by any overlayer and near surface polyimide would limit the maximum reectivity, and the observation of ca. 80% reectivity may not be limited by silver particle morphology but by absorbing surface polyimide. Many experiments with varied concentrations and cure cycles never lead to a reectivity higher than 82%. Surface metal morphology does not seem to be the main problem in limiting reectivity since commercial glass silvered mirrors have a similar surface roughness. Interestingly, TEM data for a 9.9% AgHFA

110

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

111

silvered lm showed that adjacent to the air-side-metallized surface there are several hundred nanometers which are substantially free of silver(0) particles; this depletion zone, for which we have no explanation at present, was also observed in Ag(HFA)(COD)-metallized BTDA/4,4 -ODA lms reported by Rubira et al. [50,51]. Observation of the depletion zone indicates that the in situ method of generating AgHFA is essentially equivalent to adding the solid Ag(HFA)(COD) complex, except that better reectivities were observed. The TEM micrograph clearly indicates that there is no continuity of the metal surface. SEMs of the glass side of the lms always showed less silver with more irregular particle sizes. The glass side for this 9.9% silver lm is much richer in uorine than the air side via XPS, 18.3 and 3.2%, respectively. XPS data show that the uorine is most likely due to CF3 groups with a C 1s binding energy of ca. 293 eV and a uorine binding energy of 688 eV. There is no evidence for an inorganic uoride such as that which might arise from the formation of silver(I) uoride. Thermal data for the silvered lms (Section A of Table 4) show that metallization of the polyimide does not do damage to the bulk polymer properties. Tg for all lms, regardless of concentration of silver, is very close to that of the parent. Similarly, the coefcients of linear expansion are the same as that for the parent, consistent with nanometer size silver particles which are not in contact with one another. One thermal property of the metallized lms that is degraded is the temperature at which 10% weight loss is observed. Weight loss in air for silvered lms is ca. 150 C lower than for the parent; the metallized lms are ca. 100 C more stable in a nitrogen atmosphere. The thermal stability of the silvered lms is also shown by elemental analysis data (Section A of Table 4). The silver found in the metallized lms is close to that calculated for the silver additive decomposing to silver metal and volatile ligand products which exit the lm on heating. The correspondence of calculated and found silver clearly indicates no massive polyimide oxidative degradation as might be expected. As mentioned above, when a BTDA/4,4 -ODA-AgHFA resin is cast and thermally cured on a glass plate, the metallizing lm begins to lift uniformly from glass plate in the temperature regime where silver(I) is visually seen to undergo reduction (ca. 180200 C). Indeed, the TGA curve for the 1,5-cyclooctadiene adduct of (hexauoroacetylacetonato)silver(I) [60,61] shows early loss of COD followed by reductive decomposition of AgHFA over the temperature range 160200 C. This is the same range where the lm inates and lifts from the plate. The lm remains rmly adhered at the edges, and thus the curing lm forms a dome-like structure above the glass plate. The detachment of the lm from the glass plate is presumably due to two effects. First, gas evolution arising from the release of residual solvent and/or gaseous products creates pressure on the glass side of the lm. Normally, such gases permeate through an undoped BTDA/4,4 -ODA polymer which adheres rmly to the glass plate throughout the entire cure cycle. This adhesion is

presumably due to the fact that the neat BTDA/4,4 -ODA polymer bonds strongly to the glass surface via reaction in part with surface SiOH entities. Thus, no bubbling is observed for silver-free lms. However, and secondly, the formation of silver metal at the glasspolymer interface eliminates the strong adhesion between the glass plate and the curing polymer since silver(0) is a passive and non-oxophilic metal. Thus, there is little tendency for silver to adhere to the glass via AgOSi bonds, and the curing polymer lm rises from the glass plate with the increasing gas pressure since the polymer is still very exible at 200 C as the partially imidized poly(amic acid) which is plasticized with residual solvent and silver(I) additive. Eventually, at temperatures beginning near 225 C the lm relaxes back onto the glass plate, so that if one had not observed the lm having been inated-off the plate during the cure cycle, one would not have known that the lm had been anything but resting at against the plate during the entire cure cycle. This detachment is a serious problem for coatings which are to be cured in place. Detachment is not a problem for lms which are ca. 225 cm2 in area or smaller and are to be used apart from the substrate on which they were cured. However, lms with larger areas lift from the plate in sections which leaves crease lines. This detachment problem was been solved by Southward et al. by simply casting silver(I)-BTDA/4,4 -ODA lms on undoped, fully imidized parent BTDA/4,4 -ODA lm bases. Data for representative lm-on-lm composite lms are shown in Section C of Table 4. These composite lms remain rmly adhered to the glass base and give smooth and uniformly at-metallized panels. Furthermore, there is outstanding adhesion both at the polymerpolymer interface and at the metalpolymer interface. The lms have strongly adhered surface metal layers and are completely stable to removal of silver by a variety of adhesive tapes as per the ASTM testing protocol. The polymerpolymer interface never exhibited any signs of separating even with soaking in water for months. While fully imidized BTDA/4,4 -ODA is generally insoluble in DMAc and other organic solvents, it may be that the strong polymerpolymer interface is due to the fact that on standing for 18 h at room temperature before being subjected to thermal curing a very small portion of the doped BTDA/4,4 -ODA lm dissolves into the base lm to give excellent adhesion, perhaps involving transimidization between the two layers. Since Tg for BTDA/4,4 -ODA is ca. 275 C, and the composite-metallized lms are cured some 25 C above Tg , the two polymer lms may also undergo thermal welding. Studies by Kramer et al. [34] on PMDA/4,4 -ODA, which has a Tg of 380 C, showed that a second PMDA/4,4 -ODA lm cast on a fully imidized PMDA/4,4 -ODA base lm gave a very sharp interfacial boundary, and the layers were easily peeled apart. They concluded that strong interfacial bonding occurred only when an inter-penetration layer of more than 50 nm occurred; with a diffusional distance of 200 nm the composite lm exhibited the same strength as the bulk material. TEM

112

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

data, presented later in this paper, show a distinct boundary between the silver-containing BTDA/4,4 -ODA layer and the BTDA/4,4 -ODA base layer. Thus, it appears that there is minimal diffusion of silver atoms and clusters into the base BTDA/4,4 -ODA lm. The mechanical and thermal characteristics for the lm-on-lm samples are essentially those of the parent polyimide. This lm-on-lm approach minimizes the silver required for the formation of a reective surface. Sawada and Ando [62] have been interested in the synthesis and properties of polyimides derived from 2,2 -bis(triuoromethyl)-4,4 -diaminobiphenyl (TFDB) because of their potential for optical communication applications. For example, the polyimides derived from the dianhydrides PMDA or 6FDA and TFDB as the diamine shows high transparency over 400800 nm (visible) and low optical transmission losses in the near-infrared region, as well as low a dielectric constants, low refractive indices, and low water absorption. Waveguides constructed of these materials have optical losses of less than 0.3 dB/cm at 1300 nm, and the increase in optical loss was less than 5% after heating at 380 C for 1 h and at 85 C with a relative humidity of 85% for over 200 h. The incorporation of nanometer-sized metal particles into uorinated polyimides to effect property modication was pursued. Silver(I) acetylacetonate with hexauoroacetylacetone was added to the poly(amic acid) of PMDA/TFDB to give a homogeneous solution from which lms were cast. The HFAH was added to improve the solubility of the metallic dopant. No quantitative data were given with regard to the amount of HFAH added, and thus their experiment cannot be duplicated from their published report. Thus, the active dopant species is (hexauoroacetylacetonato)silver(I) as the acetylacetonate ligand would be exchanged via proton transfer from the much more acidic HFAH to the basic acetylacetonate ligand. PMDA/TFDB lms were cast as doped poly(amic acid) solutions and cured on fused silica. All lms were cured under nitrogen from 70 to either 300 or 350 C at a heating rate of 4 K/min. (This is rapid compared with previously described work.) Ten percent weight loss was done only under nitrogen and was ca. 100 C less than that of the parent polymer. Glass transition temperatures were not determined. The silvered lm was said to be fragile although no mechanical measurements were reported. After curing to 300 C the lm appeared reddish brown on both the air and silica sides. However, after curing to 350 C a lustrous metallic surface appeared. The lms were not conductive, and no specular reectivity measurements were made. The goal appeared to be the preparation of lms in which there is a uniform distribution of nanometer-sized metal particles evenly dispersed throughout the lm. X-ray diffraction patterns for lms cured to 300 and 350 C were sharp and indicated only f.c.c. silver metal. From the Scherrer equation the crystallite size of the 350 C lms was ca. 17 nm. No peaks for silver(I) oxide or silver(I) compounds were observed. TEM data showed

silver particle sizes from 10 to 30 nm uniformly dispersed throughout the lm. It was especially interesting to note that at the air side surface of the 350 C lm there was an ca. 30 nm band of silver particles ca. 50 nm below the lm surface. Such an overlayer has been suggested the work of Taylor and coworkers and Southward et al. explaining the lack of surface conductivity and the decreasing reectivities of non-conducting lms with increasing angle of incidence. The work presented to this point with an HFA complex of silver(I) was encouraging with regard to preparing silvered lms which had metallic-like specular reectivity, but discouraging with regard to preparing lms with even modest surface conductivity. However, on making a subtle change to from hexauoroacetylacetone to triuoroacetylacetone as the -diketonate ligand source dramatic metallization results were eventually achieved as described next. 2.7. (1,1,1-Triuoroacetylacetonato)silver(I) studies In 1997, Southward et al. [63,64] reported the initial metallization studies with (triuoroacetylacetonato)silver(I), AgTFA, prepared via the in situ approach used for AgHFA. AgTFA was rst isolated in 1981 by Wenzel and Siever [65]. However, it proved extremely difcult to prepare in a reproducible manner and is photolytically unstable. Thus, with AgTFA an in situ preparation for each individual lm preparation was invaluable, assuring that lm variation was not a function of degraded complex. As selected data in Table 5 show, the rst AgTFA lms had only modest specular reectivity (2952%) relative to the AgHFA analogs discussed earlier for which specular reectivities at similar silver concentrations were in the range 7582%. All AgTFA lms cured only to 300 C for 1 h had a metallic green luster and showed reectivities at 20 C in 3035% range. However, while reectivities were disappointing, for the rst time consistent surface conductivity was obtained. Conductivity for AgTFA lms is a function of silver concentration and the thermal cure cycle; only at higher silver concentrations, ca. 12% or higher, and higher nal cure or post-cure temperatures of 340 C do the lms become conductive as taken from the oven. (None of the earlier lms was cured for longer than 1 h at the nal temperature of 300 or 340 C.) They also exhibited an increase in specular reectivity, which now did not diminish as a function of increasing angle of incidence. This is to be contrasted with the AgHFA analogs for which surface conductivity was never observed even with heating for 7 h in air at 300 C; the same lack of conductivity was true for metallized lms prepared with the isolable Ag(HFA)(COD) complex. All AgTFA-metallized lms are exible and can be creased tightly without breaking. The surface silver is well adhered to the polymer as observed in analogous AgHFA systems. The linear coefcients of thermal expansion (CTE) for the 10.6, 13.3 and 12.0% AgTFA lms of Table 5 are 28, 33 and 36 ppm/K, respectively. These are lower than the CTE for the parent polymer at 43 ppm/K and lower than

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

113

114

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

those observed in the analogous AgHFA lms (Section A of Table 4) at 43, 44 and 43 ppm/K for 5.0. 7.4 and 9.9% silver, respectively. Thus, in the conductive AgTFA lms one may be seeing a hybrid value for the CTE reecting the fact the surface silver aggregates are in contact with one another; the CTE of metallic silver is 19 ppm/K. It is possible that there is silver promoted crosslinking of the polymer which is giving a lowered CTE, although we do not see any large increase in Tg values as might be expected for crosslinked lms [6668]. The TFA ligand differs only slightly from the HFA ligand by replacement of uorines in one methyl group with hydrogens. This renders the TFA anion slightly more basic (pKa = 6.3) [69] than the HFA anion (pKa = 4.35) [70]. Thus, there seemed to be little reason to expect that the Ag(I)-TFA system should give metallized lms which differed signicantly from those prepared with the HFA analog. The visual appearance of these AgTFA lms and their reectivities are strikingly different from the closely related AgHFA systems and are consistent with observations that ligand or anion effects are pronounced in self-metallizing polyimide systems. The X-ray diffraction patterns (not shown) for the 12.0% pair of lms of Table 5 reect the obvious visual difference seen in 300/340 C lm pairs. For the lms cured only to 300 C, one observes the four intense reections (1 1 1, 2 0 0, 2 2 0 and 3 1 1) of f.c.c. crystalline silver. However, these reections are broad and are consistent with the particle size of silver clusters being in the nanometer range as suggested by the Scherrer equation. After sintering at 340 C the reections become much sharper. The X-ray patterns suggest differing degrees of aggregation of silver. The TEMs of the 12.0% pair of lms show distinct differences. The lm cured to 300 C, which exhibits a metallic green sheen, has only a minimal concentration of metallic silver at the air-side surface and a collage of silver particles dispersed throughout the bulk of the polymer matrix. The TEM for the 340 C post-cured lm shows a well-dened surface layer of silver with a thickness on the order of 200 nm. On an average, the size of the silver particles in the post-cured lm are larger than those in the lm cured only to 300 C, which is consistent with the narrower X-ray reections of this lm. An SEM of the 300 C 12.0% lm shows a much smoother and less-detailed surface structure than that for the 340 C sintered lm. At 300 C the lm exhibits what appears to be isolated silver islands emerging from the surface of the polymer. The islands are not in contact, and therefore the lm is not conductive. Upon sintering at 340 C the SEM shows a substantial increase in silver at the lm surface, and the surface morphology is such that there are conductive metallic pathways extending over the polymer. In summary, these rst studies with the (triuoroacetylacetonato)silver(I) complex, prepared in situ from silver(I) acetate and triuoroacetylacetone, showed that proper control of concentration and cure temperatures gives exible silver surface-metallized lms which are conductive but exhibit only modest reectivity. The reectivity of

the conductive lms can be increased to >80% for some lms with only light polishing. The surface layer of silver is strongly adhered to the polymer, presumably held in place by a mechanical interlocking mechanism. This work made it clear that morphology and properties of in situ self-metallized silver lms are subtle phenomena which are strongly dependent on the nature of the silver(I) species that is dispersed in the initial polymer solution from which lms are cast. The question which remained was whether silvered lms could be produced which had both metallic-like conductivity and reectivity. Very recently, Southward et al. [7173] reported additional experiments with the BTDA/4,4 -ODA-AgTFA system and found that it was possible to prepare lms with both exceptional reectivity and conductivity. The development of reectivity as a function of temperature and time was examined. While reectivity develops slowly on heating to 300 C, remaining for 47 h at 300 C gives superb lms with reectivities close to that of native silver. The only difference between the initial AgTFA work and these recent efforts is that the cure cycle is different. In this latter work the doped lm is cured more slowly in the 135300 C range than before, and the lms are held at 300 C for longer periods of time rather than removing the lm from the oven after 1 h and then post-curing the lm in a subsequent step. It appears that, in addition to silver particle migration/aggregation, a major factor in producing a well-metallized surface is silver promoted surface oxidative degradation of polyimide to volatile products which then forces nanometer size silver aggregates into contact. Thus, in self-metallization experiments to this point, migration alone seems insufcient to development a conductive surface with metallic reectivity without polishing. It is also interesting to note that for the highly reective and conductive lms of Table 6 there is minimal angle dependence of the reectivity which is what one would expect of pure silver metal. This is consistent with surface degradation of the strongly absorbing surface BTDA/4,4 -ODA polyimide. Table 6 displays characterization data for three representative lms cured at 300 C for 7 h. The second entry is a 13.0% silvered lm (120 200 mm2 in area) cast from the same silver-doped resin that was used to prepare small glass slides (27 46 mm2 ) coated with metallized polymer for the reectivity versus temperature/time plot of Fig. 1 (vide infra). Entries 3 and 4 are for lms prepared from the same resin solution. The lms differ only in that one was cast directly onto a glass plate whereas the other silver-doped lm was cast more thinly onto an undoped BTDA/4,4 -ODA polyimide lm (afxed to a glass plate) previously cured to 300 C. Casting the silver-doped resin onto a parent base conserves silver and ensures that mechanical and thermal properties of these lm-on-lm composites closely resemble those of the pure polyimide as previously mentioned. Adhesion between surface silver and polyimide, and adhesion between parent polyimide and silver-doped BTDA/4,4 -ODA layer is excellent. Boiling lms in water

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

115

Table 6 Reectivity, thermal, and resistivity data for (1,1,1-triuoro-2,4-pentanedionato)silver(I)-BTDA/ODA lms cured to 300 C for 7 h and cast on a glass plate (lms 13) and on a parent polyimide base (lm 4) Film No. Percent of silver (ca.)a Parent 13.0 12.8 12.8
a b

Percent of reectivity (at 531 nm)b 20 NA 98 98 95 30 97 98 95 45 97 98 94 55 95 99 92 70 91 92 91

Tg by DSC ( C) 275 276 270 273

CTE (ppm/K) 42.8 34.3 33.0 32.8

Surface resistivity ( /sq) 7 E15 <0.1 <0.1 <0.1

Tensile strength (ksi) 19.7 19.7 20.2 19.3

Modulus (ksi) 450 474 468 445

1 2 3 4

Calculated for the AgTFA system decomposing to silver metal and volatile components which are lost from the lm. Reectivity data are relative to a Perkin-Elmer polished aluminum optical mirror set at 100%.

does not give any delamination of silver from polyimide or parent polyimide from doped polyimide. Also, no adhesive tapes removed any silver from the surface as per the ASTM adhesion testing protocol. Composite lms were cast to give a doped to undoped layer thickness ratio of 1:6. Previous work [60,61] shows that there is no diffusion of silver into the undoped BTDA/4,4 -ODA base. The specular reectivity values for the lms of Table 6 are greater than 90%. There is little angle dependence consistent with a reective surface that is dominated by silver metal [74]. The lms have sheet resistivities consistent with a metallic silver surface, when the lms reach a highly conductive and reective state. TEM

Fig. 1. Development of reectivity as a function of temperature/time for (triuoroacetylacetonato)silver(I) in BTDA/4,4 -ODA at 10.7% silver cured in air. Time zero is after holding the lm at 135 C for 1 h; from 135 to 300 C the temperature increase 0.681 /min; at 300 C the temperature remained constant. After 5 h at 300 C the lms become conductive.

data show a surface silver layer which is ca. 200 nm thick. The SEM for a AgTFA lm (Fig. 2) cured for 7 h at 300 C (%R = 95%) reveals that the air-side surface has an irregular but distinctive topography which would sustain the observed conductivity. The networked, porous, sponge-like texture of the surface is due to a combination of sintering and oxidative surface polymer degradation. It is instructive to compare this SEM with that for the closely related 10.7% AgHFA-BTDA/ODA lm cured at 300 C for 7 h where the two systems differ only by one triuoromethyl group. The surface micrograph of the non-conductive AgHFA sample displays a very different morphology. There are discrete isolated metal particles; the particles are separated from one another by intervening surface polymer. The difference between the AgTFA and AgHFA systems is striking and demonstrates that subtle changes in ligand structure have pronounced effects on lm properties. XPS data for three of the 14 lms (10.7% silver) used to generate the reectivity curve of Fig. 1 are displayed in Table 7. For the lm cured to 275 C, which is minimally reective and has little silver at the surface, XPS data for the reective surface show only 2.8% Ag and a C, N, O ratio which is close to that of the parent polyimide. For the lm cured to 300 C for 0 h the measured reectivity is 28% with an increase in surface Ag to 4.0%; the C, N, O ratio remains near that of the parent. However, on curing at 300 C for 5 h the air-side silver concentration jumps to 27% with a measured reectivity of 80%. Still, there is substantial carbon at the surface, although ion milling experiments show that further into the surface silver is 98100%. We suggest that the carbon-containing material is partially degraded surface BTDA/4,4 -ODA polymer. This is supported by the large atom percent of oxygen, which we presume is from partial oxidation of polyimide. Thus, part of the mechanism to develop a highly reective and conductive surface involves oxidative degradation of polymer. This is further supported by the fact that curing a AgTFA-BTDA/4,4 -ODA lm in nitrogen does not give a highly reective surface. The glass side of the lm cured to 300 C for 5 h has much less silver and an organic composition closer to that of the parent polymer. The glass transition temperatures of the AgTFA-metallized lms do not vary from that of undoped BTDA/4,4 -ODA

116

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

Fig. 2. SEM of selected silver-polyimide lms: (top) Ag(CF3 CO2 )-BTDA/4,4 -ODA 10.7% silvered lm cured to 300 C for 1 h (air side), R = 65% (97% polished), not conductive; (middle) AgHFA-BTDA/4,4 -ODA 10.7% silvered lm cured to 300 C for 7 h (air side), R = 62% (the maximum R was 75% after curing for 1 h at 300 C.), not conductive; (bottom) AgTFA-BTDA/4,4 -ODA 12.8% silvered lm cured on a BTDA/ODA base to 300 C for 7 h (air side), R = 95%, resistivity <0.1 /sq.

by more than a few degrees. This suggests that the bulk polymer structure is not compromised by the reduction of silver(I) and the formation of silver(0). However, the formation of metal clusters in the bulk of the polymer as well as on the surface diminishes the high temperature thermal oxidative stability of the hybrid lm. While in a nitrogen

atmosphere the temperature at which there is 10% weight loss is not vastly different from that of the undoped polymer, in air there is a reduction in stability with 10% weight loss temperature, i.e., ca. 150200 C lower than the control. Nonetheless, the thermal stability of the mirrored lms in air is more than adequate for most purposes.

Table 7 X-ray photoelectron spectroscopic surface composition for selected (1,1,1-triuoro-2,4-pentanedionato)silver(I)-BTDA/ODA (10.7% Ag) lms of Fig. 1 Wt.% of silver Parent 10.7 10.7 10.7 Thermal historya 300 275 300 300 for for for for 1h 0h 0h 5h Reectivity (%) at 20 NA 20 28 79 Resistivity ( /sq)b NA NC NC PC Film surface Air Air Air Air (glass) Ag Relative atom percent F 0 2.8 4.0 27 (3.7) 0 0.0 0.47 0.33 (4.5) C 78 78 78 45 (69) O 16 15 13 26 (19) N 5.4 4.4 4.5 1.6 (3.1)

a All cure cycles involved heating from 22 to 135 C over 20 min and holding for 1 h followed by heating over 240 min to 300 C and holding at 300 C for varying times. Samples were removed at various times in the cure. b Four-point probe. PC: partially conductive with some but not all portions of the lm registering conductivity; NC: not conductive; NA: not applicable.

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119

117

Fig. 3. Development of reectivity as a function of temperature/time for 13% silver(0) BTDA/4,4 -ODA lms with varying silver(I) additives. The silver(I) precursors are listed at the right of the plot. Time zero is after holding the lm at 135 C for 1 h; from 135 to 300 C the temperature increase 0.69 C/min; at 300 C the temperature remained constant.

The linear CTE for the three lms of Table 6 are 33 34 ppm/K. These are signicantly lower than for the parent polymer at 43 ppm/K and than for the analogous AgHFA lms at 43, 44 and 43 ppm/K for 5.0. 7.4 and 9.9% silver, respectively (Section A of Table 4). Thus, as mentioned previously, in the conductive AgTFA lms we may be seeing a hybrid value for the CTE reecting the fact the surface silver aggregates are in contact with one another. Again, the variation in CTE values between the AgTFA and AgHFA systems demonstrates the pronounced differences that can arise due to subtle ligand effects. For metallized lms of Table 6 the tensile strengths are, within experimental error, those of the parent. The modulus for the two lms cast on a glass plate are elevated ca. 5% while the value for the silvered lm cast on a BTDA/4,4 -ODA base is virtually the same as for the undoped polyimide. The AgTFA complex is introduced to the poly(amic acid) resin via an in situ synthesis. While the preparation of solid AgTFA has been reported in the literature [65], it is unstable and difcult to prepare reproducibly. Thus, we chose to synthesize AgTFA in poly(amic acid) solutions via the in situ reaction of silver(I) acetate, AgOAc, and TFAH, both of which can be obtained in pure and stable form. Combining AgOAc, TFAH, and BTDA/4,4 -ODA in DMAc under ambient conditions leads to formation of a homogeneous solution of the AgTFA complex. The basicity of the acetate anion and the large formation constants associated with metal-diketonate complexes lead to proton transfer from TFAH to the acetate anion giving AgTFA and

acetic acid. Even though AgTFA is not isolated, we can be assured that this complex is formed. First, Ag AgOAc by itself is not soluble in DMAc. Second, when the poly(amic acid) is added to AgOAc alone, immediate polymer gelation occurs. The pathway for gelation is deprotonation of the aromatic carboxyl groups of the poly(amic acid) via transfer to the more basic acetate ions of AgOAc. Carboxylate groups of the polymer then coordinate to Ag(I) ions to form the extended gel network. It is well known that silver carboxylate complexes are dimeric involving the coordination of two carboxylate groups [36,37]. If TFAH is added to AgOAc before the addition of the poly(amic acid), no gelation occurs, and the insoluble AgOAc dissolves. We chose AgOAc as the precursor to formation of the in situ AgTFA complex because the acetate salt is readily available in high purity, is thermally and photochemically stable, and is not hygroscopic. We found that TFAH/Ag(I) ratios of 1.35:1 rather than 1:1 gave doped solutions which were somewhat less viscous and more easily processed. In summary, BTDA/ODA polyimide lms can be prepared from single phase silver(I) acetate-1,1,1-triuoro-2,4-pentanedione-BTDA/ODA solutions cast and cured either on glass plates or a parent polyimide base. Depending on concentration and thermal conditions, metallized lms can be fabricated with excellent specular reectivity, surface conductivity, outstanding metalpolymer adhesion, and intact mechanical characteristics. This lm-on-lm approach minimizes the silver required for the formation of a reective surface and assures composite polymer properties which are those of the parent polyimide.

118

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119 [9] R.E. Freeland, G. Bilyou, in: Proceedings of the 43rd Congress of the International Astronautical Federation, IAF-92-0301, Washington, DC, 1992. [10] T. Lundeberg, Lancet, 1986, 1031 pp. [11] H. Liedberg, T. Lundeberg, Urol. Res. 17 (1989) 359. [12] R.J. Angelo, DuPont, USA, 1963. [13] A.L. Endrey, DuPont, USA, 1963. [14] N. Hubin, L. Noethe, Science 262 (1993) 1390. [15] J.L. Wilbur, R.J. Jackman, G.M. Whitesides, E.L. Cheung, L.K. Lee, M.G. Prentiss, Chem. Mater. 8 (1996) 1380. [16] L. Yan, X.-M. Zhao, G.M. Whitesides, J. Am. Chem. Soc. 120 (1998) 6179. [17] P.F. Green, L.L. Berger, Thin Solid Films 224 (1993) 209. [18] G. Rozovskis, J. Vinkevicius, J. Jaciauskiene, J. Adhes. Sci. Technol. 10 (1996) 399. [19] K.L. Mittal, J. Vac. Sci. Technol. 13 (1976) 19. [20] T.T. Kodas, M.J. Hampden-Smith, The Chemistry of Metal CVD, VCH, Weinheim, 1994. [21] J.T. Spencer, in: K.D. Karlin (Ed.), Progress in Inorganic Chemistry, Vol. 41, 1993, pp. 145238. [22] P.M. Jeffries, S.R. Wilson, G.S.J. Girolami, Organomet. Chem. 449 (1993) 203. [23] W. Lin, T.H. Warren, R.G. Nuzzo, G.S. Girolami, J. Am. Chem. Soc. 115 (1993) 11644. [24] Z. Yuan, N.H. Dryden, J.J. Vittal, R. Puddephatt, J. Chem. Mater. 7 (1995) 1696. [25] N.H. Dryden, J.J. Vittal, R.H. Puddephatt, Chem. Mater. 5 (1993) 765. [26] N.H. Dryden, J.J. Vittal, R.H. Puddephatt, Can. J. Chem. 72 (1994) 1605. [27] C. Xu, T.S. Corbitt, M.J. Hampden-Smith, T.T. Kodas, E.N. Duesler, J. Chem. Soc., Dalton Trans. (1994) 2841. [28] C. Xu, M.J. Hampden-Smith, T.T. Kodas, Adv. Mater. 6 (1994) 746. [29] G.M. Bower, L.W. Frost, J. Polym. Sci. A 1 (1963) 3135. [30] C.E. Scroog, A.L. Endrey, S.V. Abramo, C.E. Berr, W.M. Edwards, K.L. Olivier, J. Polym. Sci. A 3 (1965) 1373. [31] D.J. Hucknall, Selective Oxidation of Hydrocarbons, Academic Press, London, 1974. [32] L.J. Gerenser, J. Vac. Sci. Technol. A 6 (1988) 2897. [33] L.J. Gerenser, J. Vac. Sci. Technol. A 8 (1990) 3682. [34] L.J. Gerenser, K.E. Goppert-Berarducci, R.C. Baetzold, J.M. Pochan, J. Chem. Phys. 95 (1991) 4641. [35] L.J. Gerenser, K.E. Goppert-Berarducci, in: K.L. Mittal (Ed.), Metallized Plastics, Vol. 3, Plenum Press, New York, 1992, pp. 163178. [36] A.E. Blakeslee, J.L. Hoard, J. Am. Chem. Soc. 78 (1965) 3029. [37] P. Coggin, A.T. McPhail, J. Chem. Soc., Chem. Commun. (1972) 91. [38] A.E. Martell, Stability Constants of Metal Ion Complexes, Special Publication 17, The Chemical Society, Burlington House, London, 1964. [39] A.K. St. Clair, L.T. Taylor, J. Appl. Polym. Sci. 28 (1983) 23932400. [40] J.D. Rancourt, G.M. Porta, L.T. Taylor, in: Nations Future Materials Needs, Proceedings of the International SAMPE Technical Conference, Vol. 19, 1987, pp. 564575. [41] L.T. Taylor, in: W.D. Weber, M.R. Gupta (Eds.), Recent Advances in Polyimide Science and Technology, Mid-Hudson Chapter SPE, New York, 1987, pp. 428437. [42] R.K. Boggess, L.T. Taylor, in: W.D. Weber, M.R. Gupta (Eds.), Recent Advances in Polyimide Science and Technology, Mid-Hudson Chapter SPE, New York, 1987, pp. 463470. [43] M.A. Linehan, D.M. Stoakley, A.K. St. Clair, in: Proceedings of the 44th Southeastern26th Middle Atlantic Combined Regional Meeting of the American Chemical Society, Abstracts of Papers, American Chemical Society, Washington, DC, 1992, pp. POLY 378. [44] L.T. Taylor, A.K.St. Clair, in: K.L. Wittal (Ed.), Polyimides: Synthesis, Characterization and Applications, Vol. 2, Plenum Press New York, 1984, pp. 617645. [45] A. Auerbach, J. Electrochem. Soc. (1984) 937.

3. General conclusions It now appears certain any silver(I) compound which is soluble in a solution of a poly(amic acid) will lead to the incorporation of metallic silver in a polyimide lm after an appropriate thermal treatment. However, there are pronounced ligand and anion effects on the reduction of silver(I) and subsequent aggregation of silver(0) particles, and therefore, the quality of the silver surface with regard to specular reectivity and electrical conductivity varies enormously. These effects are illustrated in part in Figs. 2 and 3. The SEM micrograph of Fig. 2 shows the prominent effect that ligand variation has on silver particle morphology at the reective air side surface of the lm. Fig. 3 shows the emergence of the air side reective surface (in percent specular reectivity) as a function of cure time and temperature. It is apparent from the curves that the triuoroacetato and hexauoroacetylacetonato complexes of silver(I) metallize much earlier than the triuoroacetylacetonato complex, and that the former two additives lead to larger discrete silver particles at the surface (Fig. 2). However, these larger particles are isolated from one another by the dielectric polyimide and hence are not conductive. The (triuoroacetylacetonato)silver(I) complex with BTDA/4,4 -ODA gives much smaller particles throughout most of the cure cycle, which is clear from TEM micrographs and line broader X-ray diffraction patterns which are not shown [72]. These smaller silver(0) particles appear to catalyze surface polyimide degradation after the lm is at 300 C for ca. 23 h. This polymer degradation process allows metal particles to come into contact giving a conductive surface. Figs. 2 and 3 are simply meant to reinforce the point that metallization effects are not well understood. The mechanism of silver(I) reduction and factors affecting aggregation and migration remain unclear; the roles of ligand, solvent, and polymer functionality in the reduction and aggregation processes need systematic study. Nonetheless, synthetic pathways now exist to fabricate-metallized lms and coatings from passive metals.

References
[1] R. Gliem, G. Schlamp, Metall. 41 (1987) 34. [2] P.A. Gierow, in: Proceedings of the ASMEJSMEJSES Solar Energy Conference, Reno, NV, 1991, pp. 17. [3] K. Ehricke, in: Meeting of the American Rocket Society, Cleveland, OH, June 1820, 1956, ARS Paper, pp. 310356. [4] D.A. Gulino, R.A. Egger, W.F. Bauholzer, NASA Technical Memorandum 88865, 1986. [5] V.B. Hueggle, in: Reective Optics II, Proceedings of Society of Photo-optical Instrumentation Engineers, Vol. 79, 1989, 1113 pp. [6] H.H. Neidlinger, P. Schissel (Eds.), Polymers in Solar Technologies, VCH, Weinheim, 1988, pp. 3451. [7] H.H. Neidlinger, P. Schissel, in: Proceedings of Society of Photooptical Instrumentation Engineers, Vol. 823, 1987, 181 pp. [8] G. Jorgensen, P. Schissel, in: K.L. Mittal, J.R. Susko (Eds.), Metallized Plastics, Vol. 1, Plenum Press, New York, 1989, pp. 7992.

R.E. Southward, D.M. Stoakley / Progress in Organic Coatings 41 (2001) 99119 [46] A. Auerbach, J. Electrochem. Soc. (1985) 132. [47] R.E. Southward, D.S. Thompson, D.W. Thompson, M.L. Caplan, A.K. St. Clair, Chem. Mater. 7 (1995) 21712180. [48] R.E. Southward, C.M. Boggs, D.W. Thompson, A.K. St. Clair, Chem. Mater. 10 (1998) 14081421. [49] R.E. Southward, D.W. Thompson, S.H. Sproul, J.L. Wasyk, J.L. Scott, S.T. Broadwater, C.M. Boggs, A.K. St. Clair, Polym. Mater. Sci. Eng. 78 (1998) 1314. [50] A.F. Rubira, J.D. Rancourt, M.L. Caplan, A.K. St. Clair, L.T. Taylor, Polym. Mater. Sci. Eng. 71 (1994) 509. [51] A.F. Rubira, J.D. Rancourt, M.L. Caplan, A.K. St. Clair, L.T. Taylor, Chem. Mater. 6 (1994) 2351. [52] A.F. Rubira, J.D. Rancourt, M.L. Caplan, A.K. St. Clair, L.T. Taylor, in: C.U. Pittman, C.E. Carraher, B.M. Culbertson, M. Zeldin, J.E. Sheets (Eds.), Metal-containing Polymeric Materials, Plenum Press, New York, 1996, pp. 357366. [53] A.F. Rubira, J.D. Rancourt, M.L. Caplan, A.K. St. Clair, L.T. Taylor, in: C. Feger, M.M. Khojasteh, S.E. Molis (Eds.), Polyimides: Trends in Materials and Applications, Society of Plastics Engineers, Mid-Hudson Section, New York, 1996, 475 pp. [54] A.F. Rubira, J.D. Rancourt, L.T. Taylor, D.M. Stoakley, A.K. St. Clair, J. Macromol. Sci. Pure Appl. Chem. A 35 (1998) 621636. [55] A. Bailey, T.S. Corbitt, M.J. Hampden-Smith, E.N. Duesler, T.T. Kodas, Polyhedron 12 (1993) 1785. [56] W. Partenheimer, E.H. Johnson, Inorg. Chem. 11 (1972) 2840. [57] M.L. Caplan, R.E. Southward, D.W. Thompson, A.K. St. Clair, Polym. Mater. Sci. Eng. 71 (1994) 787788. [58] R.E. Southward, D.S. Thompson, D.W. Thompson, M.L. Caplan, A.K. St. Clair, Polym. Mater. Sci. Eng. 73 (1995) 382383.

119

[59] R.E. Southward, D.S. Thompson, D.W. Thompson, M.L. Caplan, A.K. St. Clair, in: C.U. Pittman, C.E. Carraher, B.M. Culbertson, M. Zeldin, J.E. Sheets (Eds.), Metal-containing Polymeric Materials, Plenum Press, New York, 1996, pp. 349356. [60] R.E. Southward, D.S. Thompson, D.W. Thompson, A.K. St. Clair, Polym. Mater. Sci. Eng. 76 (1997) 185186. [61] R.E. Southward, D.S. Thompson, D.W. Thompson, A.K. St. Clair, Chem. Mater. 9 (1997) 16911699. [62] T. Sawada, S. Ando, Chem. Mater. (1998) 10. [63] R.E. Southward, D.W. Thompson, A.K. St. Clair, Chem. Mater. 9 (1997) 501509. [64] R.E. Southward, D.W. Thompson, A.K. St. Clair, Polym. Mater. Sci. Eng. 74 (1996) 414415. [65] T.J. Wenzel, R.E. Sievers, Anal. Chem. 53 (1981) 393. [66] J.H. Glans, D.T. Turner, Polymer 22 (1981) 1540. [67] K.-M. Park, I.-W. Shim, J. Appl. Polym. Sci. 42 (1991) 1361. [68] I.-W. Shim, W.-S. Oh, H.-C. Jeong, W.-K. Seok, Macromolecules 29 (1996) 1099. [69] R.W. Taft, E.H. Cook, J. Am. Chem. Soc. 81 (1959) 46. [70] J.C. Reid, M. Calvin, J. Am. Chem. Soc. 72 (1950) 2948. [71] R.E. Southward, D.S. Thompson, D.W. Thompson, J.L. Scott, S.T. Broadwater, A.K. St. Clair, in: Materials in Space: Science, Technology, and Applications 551 (1999) 453459. [72] R.E. Southward, D.S. Thompson, D.W. Thompson, A.K. St. Clair, Chem. Mater. 11 (1999) 501507. [73] R.E. Southward, D.W. Thompson, A.K. St. Clair, Polym. Prepr. 39 (1) (1998) 423424. [74] L.F. Drummer, G. Haas, in: G. Haas, R.E. Thun (Eds.), Physics of Thin Films, Vol. 2, Academic Press, New York, 1964, pp. 305361.

S-ar putea să vă placă și