You are on page 1of 17

Downloaded from rsta.royalsocietypublishing.

org on September 10, 2012

Control of nucleation in glass ceramics


Wolfram Hland, Volker Rheinberger and Marcel Schweiger Phil. Trans. R. Soc. Lond. A 2003 361, 575-589 doi: 10.1098/rsta.2002.1152

Email alerting service

Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand corner of the article or click here

To subscribe to Phil. Trans. R. Soc. Lond. A go to: http://rsta.royalsocietypublishing.org/subscriptions

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

10.1098/rsta.2002.1152

Control of nucleation in glass ceramics


l a n d, V o l k e r Rheinberger By W o l f r a m H o a n d M a r c e l Schweiger Ivoclar Vivadent AG, Bendererstrasse 2, 9494 Schaan, Principality of Liechtenstein
Published online 29 January 2003

Glass ceramics are advanced materials composed of one or more glass and crystal phases. By developing base glasses with appropriate compositions and by controlling crystal nucleation and growth in these glasses, glass ceramics with tailor-made properties can be fabricated. The key to developing this type of material is control of the nucleation processes. Both volume and surface nucleation can be exploited. Heterogeneous volume nucleation has been used to develop glass ceramics showing minimal thermal expansion and high strength. Two nucleation mechanisms can be combined and the precipitation of two crystal phases can be controlled. That the nucleation processes can be controlled by nano- and microscale immiscibility is a special feature, allowing selective nanophase formation or the development of needle-like apatite phases demonstrating a natural morphology. This represents a biomimetic process. The control of nucleation has enabled the development of biomaterials for dental applications.
Keywords: glass ceramics; twofold nucleation; biomimetic process; dental materials

1. Introduction
Glass ceramics are modern materials with an inorganicinorganic microstructure. This dense and void-free microstructure consists of one or more glassy and crystalline phases. The chemical composition and microstructure of the glass ceramic determine its properties and main applications. Glass ceramics are most commonly made by forming special base glasses, mostly by melting, and then using controlled heat treatment to nucleate and precipitate crystals in the glassy matrix. H oland & Beall (2002) have reviewed the formation of glass ceramics based on alkaline and alkaline earth silicates, aluminosilicates, uorosilicates, silicophosphates, iron silicates, phosphates, niobates or titanates, with special properties for specic applications. The microstructures that can be formed include nanoscale phases, highly uniform crystals with an interlocking microstructure, and crystals similar to those in natural bone.

2. Mechanisms of controlled nucleation


Nucleation is the key factor for controlling crystallization in glass ceramics. Classical theory can describe the temperature dependence of nucleation and crystallization in
One contribution of 15 to a Discussion Meeting Nucleation control. Phil. Trans. R. Soc. Lond. A (2003) 361, 575589 c 2003 The Royal Society

575

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

576

W. H oland and others

glasses. The nucleation rate and crystal growth rate as a function of temperature are accurately measured experimentally (Stookey 1959; McMillan 1979, H oland & Beall 2002). Nucleation mechanisms have been reviewed by Kingery et al . (1975), Zanotto (1994), James (1982) and Weinberg et al . (1997). Two general nucleation mechanisms have been exploited to develop glass ceramics: volume nucleation and surface nucleation. In this article, these two mechanisms will be discussed from the point of view of practical applications. (a ) Volume nucleation (i) Homogeneous and heterogeneous nucleation In the crystallization of glasses, as in other phase transformations, a distinction must be made between homogeneous and heterogeneous nucleation. For homogeneous nucleation, the classical theory gives the work of formation G of a spherical nucleus of radius r as 3 2 (2.1) G = 4 3 r Gv + 4r + GE , where Gv is the free-energy change per unit volume associated with the formation of the new phase, is the interfacial energy (per unit area) of the new surface of the nucleus, and GE is the elastic distortion energy (often not considered). To develop glass ceramics, special nucleation agents can be incorporated into the base glass that acts as a catalyst for nucleation in the glassy matrix. For nucleation on catalytic substrates, the classical picture is of a spherical-cap nucleus making a contact angle with the substrate. The critical work of formation for heterogeneous nucleation, G H , is lower for smaller according to
G H = G f ( ), 2 where f () = 1 4 (2 + cos )(1 cos ) .

(2.2)

Based on (2.2), the nucleation barrier is very small if the nucleant substrate is completely wetted by the nucleus and the contact angle is close to 0 . Heterogeneous nucleation is particularly eective if there is epitaxy between the nucleus and substrate. There can be an epitaxial relationship if the lattice geometry of the nucleus and substrate crystals is similar (less than 15% mismatch in lattice parameter). Further inuences on epitaxy in glass ceramics include: the bonding state in the substrate and nucleus crystals, structure defects, and the degree of coverage of the nucleant surface with foreign nuclei. Weinberg et al . (1997) and Zanotto (1994) have reported in detail on crystallization kinetics in glasses. The standard theory of this type of phase-transformation kinetics was developed by Johnson and Mehl, Avrami and Kolmogorov (JMAK theory). The kinetics of the nucleation rate, I , were investigated by James (1982) and Gutzow (1980) as functions of time t and one suggested form is I = I0 exp , t (2.3)

where I0 is the steady-state nucleation rate and the non-steady-state time lag, a time before the steady-state rate is reached. The steady-state rate can be expressed as G + G D , (2.4) I0 = A exp kT
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics

577

where G is the work of formation of the critical nucleus and is the thermodynamic barrier, GD represents the kinetic barrier for nucleation, k is Boltzmanns constant, and A a pre-exponential factor. I0 is inversely proportional to the viscosity, , of the glass phase. The characterization of the nucleation behaviour is possible by experimental determination of the nucleation rate, I , and the nucleus number, N , as a function of the time of heat treatment under isothermal conditions. Thus, the functions I (t) and N (t) can be determined at dierent temperatures. The time lag can also be characterized. From the nucleation kinetics it may be possible to distinguish the mechanism. In many cases of glass-ceramic formation, the crystalline phases are metastable. Temperaturetimetransformation (TTT) diagrams are convenient for displaying the competition between metastable and stable phases for a given glass-ceramic composition (Uhlmann 1980). TTT diagrams for multi-component systems have been shown by H oland & Beall (2002). (ii) The role of micro-immiscibility Particular glasses from the systems SiO2 Al2 O3 Li2 O, SiO2 Al2 O3 MgO and SiO2 Al2 O3 K2 O tend to phase separate into a SiO2 -rich glassy matrix and an alkaline or alkaline-earth oxide-rich glassy droplet phase. The relationship between nucleation and micro-immiscibility was summarized by Uhlmann & Kolbeck (1976). First, by using phase separation in the base glass, volume crystallization can be achieved at an earlier stage or delayed by changing the composition of the matrix phase; surface crystallization or uncontrolled volume crystallization can thus be suppressed. Second, phase separation may lead to the formation of a low-viscosity phase demonstrating homogeneous crystallization, while the matrix crystallizes heterogeneously, either simultaneously or later. Third phase separation leads to the formation of interfaces that may be preferred sites for crystallization. Importantly, it has been clearly demonstrated that phase separation can be controlled, in turn enabling further control of the nucleation of primary crystals. With phase separation, material transport delivers the molecular building blocks for the nucleus to the nucleation site (the droplet glass phase). The term GD in (2.4) appears to be reduced to allow direct and rapid formation of the desired primary nucleus in the droplet phase. Beall & Duke (1983) showed that TiO2 , ZrO2 , P2 O5 , Ta2 O5 , WO3 , Fe2 O3 and uoride promote special microphase separation in various chemical systems permitting the development of glass ceramics. These nucleating agents may accumulate in a specic microphase of the phase-separated base glass. (b ) Surface nucleation Most glass ceramics exploit volume nucleation, but there are also base glasses in which controlled volume nucleation cannot be initiated. In these glasses, controlled crystallization can be achieved only with surface nucleation. This process, however, is more dicult to control. It is clear that nucleation and crystallization can be accelerated and controlled by tribochemical activation of the glass surface. The possibilities of controlling the nucleation process were successfully studied by the TC7 group of the International Commission on Glass (Pannhorst 2000). Their studies of glass-ceramic formation from near-stoichiometric cordierite glass showed that seeding of the glass initiates surface nucleation in addition to tribochemical reactions.
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

578

W. H oland and others

Seeding of the glass surface was possible with ne powdered glass of the same composition or heterogeneous particles, e.g. Al2 O3 . Further investigations of the kinetics of nucleation demonstrated that the elastic-strain term (GE ) in (2.1) could also be of signicance in surface nucleation. Kokubo (1991) developed bioactive glass ceramics for bone replacement in human medicine. He used the mechanism of surface nucleation to precipitate apatite and wollasonite as main crystal phases of the glass ceramic.

3. Application of nucleation control to multi-component glass ceramics


Base glasses with simple, stoichiometric compositions have been used mainly to make quantitative tests of nucleation theories. For glass ceramics with important properties for technical, consumer, medical or biological applications, however, it is necessary to develop multi-component base glasses and still to apply the dierent nucleation mechanisms to convert the base glasses into glass ceramics. (a ) Glass ceramics with low thermal expansion The development of glass ceramics having a very low or even zero coecient of thermal expansion (CTE) in a wide temperature range demonstrates the control of nucleation by phase separation of the base glass and additional heterogeneous nucleation. Glass ceramics derived from the SiO2 Al2 O3 Li2 O system were selected. The unusual CTE was achieved by forming crystals of metastable solid solutions based on -quartz and -spodumene. The key to controlled crystallization of the -quartz solid solution is controlled nucleation, by choosing the correct nucleating agent, the addition level of the agent and the heat treatment of the base glass. Beall et al . (1967) developed an eective nucleation initiation method for the precipitation of very small crystals of the -quartz solid solution in the SiO2 Al2 O3 Li2 OMgO ZnO system through additions of TiO2 , ZrO2 , or occasionally Ta2 O5 . The desired crystallites in this case were less than 100 nm in diameter and are an example of nanophase formation. The combination of the two nucleating agents TiO2 and ZrO2 was examined by Beall et al . (1967) and Petzoldt (1967). An eective nucleating agent ratio of 2 wt% TiO2 and 2 wt% ZrO2 was established (Beall 1992) for the nucleation of highly dispersed ZrTiO4 crystallites at 780 C. During subsequent heat treatment at 980 C, crystallites of -quartz-solid solution less than 100 nm across grew heterogeneously as the main crystal phase on the primary nuclei of ZrTiO4 . Given the special properties of low CTE and thermal conductivity, the glass ceramic has been successfully used as telescope mirror blanks in precision optics and for a variety of household applications. (b ) Glass ceramics with high mechanical strength The development of glass ceramics with high mechanical strength is also based on control of volume nucleation by phase separation of the base glass. High-strength glass ceramics with very good processibility (permitting pressing to shape at less than 1000 C) were developed in base glasses of SiO2 Li2 OP2 O5 ZrO2 (ZrO2 -containing oland & Beall glass ceramics), and SiO2 Li2 O (lithium disilicate glass ceramics) (H 2002).
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics

579

1 m
Figure 1. Microstructure of a ZrO2 -containing base glass with ne primary Li3 PO4 crystals. Scanning electron micrograph (SEM) of etched sample (2.5% HF, 10 s).

(i) ZrO2 -containing glass ceramics oland Glass ceramics containing ZrO2 are known in dierent chemical systems (H & Beall 2002). In the SiO2 Li2 OZrO2 P2 O5 system, glass ceramics are found in the composition range of 4259 SiO2 , 715 Li2 O, 415 P2 O5 , 1528 ZrO2 (all compositions given as wt% unless stated otherwise) with additions of K2 O, Na2 O, Al2 O3 and F up to 11 wt%. These glass ceramics can be shaped under hot-pressing conditions. In a base glass containing 20 wt% ZrO2 , nucleation is initiated by phase separation and Li3 PO4 primary crystal formation during quenching of the glass melt. Li3 PO4 crystals are visible in gure 1 on the edges of the spherical etched areas. The subsequent volume crystallization involves precipitation of ZrO2 . Two ZrO2 phases (baddeleyite-type and the tetragonal polymorph) are formed as microcrystals measuring 200300 nm at 940 C and grow up to 120 m in length during extended heat treatments. The growth rate of the ZrO2 microcrystals has its maximum at oland et al . 1996). The corresponding glass ceramics are produced 3.55.0 m h1 (H under viscous ow conditions; they show bending strengths of 280 MPa and a fracture toughness KIC of 2.0 MPa m1/2 . (ii) Lithium disilicate glass ceramics James (1982) and Deubener et al . (1993) investigated non-steady-state nucleation in lithium disilicate glass ceramics. Headley & Loehman (1984) added P2 O5 to SiO2 Li2 O glasses and discovered heterogeneous nucleation of lithium disilicate by epitaxial growth on Li3 PO4 crystals in a special glass ceramic. Beall (1993) developed chemically durable lithium disilicate glass ceramics in a multi-component system. High-strength glass ceramics were developed by Schweiger et al . (1998), based on a powder-processed multi-component lithium disilicate glass ceramic with composition 5780 SiO2 , 05 Al2 O3 , 0.16 La2 O3 , 05 MgO, 08 ZnO, 013 K2 O, 1119 Li2 O, 0.5 11 P2 O5 , 06 additives, and colouring substances 08. The powdered base glass was
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

580

W. H oland and others

5 m
Figure 2. Microstructure of a lithium disilicate glass ceramic after hot pressing at 920 C. SEM, etched sample (HF vapour, 10 s). Table 1. Crystal-phase formation in multi-component (a) Al2 O3 -free and (b) Al2 O3 -containing lithium disilicate glass ceramics (a) temp. 500 C 630 C 650 C crystal phase Li2 Si2 O5 Li2 SiO3 Li2 Si2 O5 Li2 SiO3 , SiO2 Li2 Si2 O5 Li3 PO4 temp. 580 C 760 C (b) crystal phase Li2 SiO3 Li2 Si2 O5 Li3 PO4

heat treated at 500800 C, and a raw glass ceramic was produced in the form of a cylindrical ingot. This ingot was transformed into a state of viscosity 105 106 Pa s in a special hot-press apparatus (EP 500, EP 600, Ivoclar Vivadent AG, Liechtenstein), then pressed at 920 C for 515 min to form a glass-ceramic body. This lithium disilicate glass ceramic does not require additional heat treatment. In the end product the main crystal phases are Li2 Si2 O5 and Li3 PO4 . Based on intensive studies of the predominant volume crystallization, nucleation must be initiated by P2 O5 . Schweiger et al . (1998) and H oland et al . (2000a) showed that the chemical composition has an important inuence on the nucleation and crystallization of lithium disilicate, the main crystal phase. Complex simultaneous and sequential solid-state reactions were studied in an Al2 O3 -free lithium disilicate glass ceramic of comoland et al . position (mol.%) 63.2 SiO2 , 29.1 Li2 O, 2.9 K2 O, 3.3 ZnO, 1.5 P2 O5 (H 2000a). Li3 PO4 crystals were formed after the crystallization of Li2 SiO3 and Li2 Si2 O5 (table 1a). Therefore, Li3 PO4 does not nucleate lithium disilicate crystals, as conPhil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics

581

2 m
Figure 3. SEM showing primary crystals of leucite formed by heat treatment (12 h at 720 C) of a seeded monolithic glass surface. (Reproduced from Hoeland et al . (1995) with permission of Elsevier Science.)

cluded previously for glasses of similar composition (Holland et al . 1998). Therefore, the nucleation mechanism may be based on steep compositional gradients. The crystals of Li2 Si2 O5 were precipitated at low temperatures (ca. 520 C) in parallel with Li2 SiO3 (table 1a). The growth rate of lithium disilicate increased at 680 C after the dissolution of Li2 SiO3 and SiO2 (cristobalite intermediate phase). Thus, the solidstate reaction Li2 SiO3 + SiO2 Li2 Si2 O5 may have taken place. Schweiger et al . (2000) and H oland et al . (2000b) found that Al2 O3 -containing and La2 O3 -containing multi-component lithium disilicate glass ceramics show dierent nucleation behaviour (table 1b). The crystallization of the Li2 Si2 O5 as the main crystal phase clearly occurs via the precursor phase lithium metasilicate, Li2 SiO3 , as shown by thermal analysis (H oland & Beall 2002) and X-ray diraction. The resulting microstructure has a crystallinity of (70 5) vol.% and is shown in gure 2. This glass ceramic is a translucent, high-strength material with a bending strength of 300400 MPa; it is used as biomaterial in restorative dentistry. (c ) Surface nucleation and sintering Glass ceramics derived from the SiO2 Al2 O3 K2 O system involve surface nucleation and crystallization. The main crystal phase is leucite (K Al Si2 O6 , i.e. K2 O Al2 O3 4SiO2 ). The glass ceramics are characterized by good optical properties, high CTE and good sinterability; they are also used as biomaterials in restorative dentistry. A typical glass ceramic with leucite as the main crystal phase has a composition in the range 5963 SiO2 , 1923.5 Al2 O3 , 1014 K2 O, 3.56.5 Na2 O, 01 B2 O3 , 01 CeO2 , 0.53 CaO, 01.5 BaO and 00.5 TiO2 (H oland & Beall 2002). The nucleation mechanism has been studied in a base glass of composition 63.0 SiO2 ,
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

582

W. H oland and others

2 m
Figure 4. Dendritic leucite growth in glass granulates shown on an SEM after HF etching. (Reproduced from Hoeland et al . (1995) with permission of Elsevier Science.)

17.7 Al2 O3 , 11.2 K2 O, 4.6 Na2 O, 0.6 B2 O3 , 0.4 CeO2 , 1.6 CaO, 0.7 BaO and 0.2 TiO2 (H oland et al . 1995). Three experiments were carried out: nucleation with almost at polished monolithic samples; nucleation at the at surface using seeding particles; and nucleation and crystallization of glass powders. When the glass was heat treated in monolithic form without seeding particles, the nucleation rate was low and the growth rate was 5 nm min1 . The leucite crystals demonstrate anisotropic growth in which the crystals grow rectangular to the surface of the glass specimen. Using glass dust of the same composition, nucleation in the monolithic glass led to twodimensional crystal growth (gure 3), in which a at cell-like crystal grew on the surface of the base glass at 720 C. Subsequently, this crystal grew and transformed into the predominant crystal-phase leucite. In glass powders there was heterogeneous surface nucleation of the grains by glass dust. While the leucite crystals grew on seeded areas as well as on unseeded areas (H oland & Beall 2002), seeding did considerably increase the nucleation rate and the growth rate was 2 m min1 . The population density of the nuclei can be increased and eective surface crystallization ensured by nely grinding the base glass. Applying this method to leucitecontaining glass ceramics can, on heat treating between 920 C and 1200 C, lead to petal-like, dendritic crystals (gure 4). The glass ceramic is used for dental restoration; therefore, it is important that it be translucent, even though the crystals are 25 m in diameter. The exural strength of the nal product is 120140 MPa. Furthermore, leucite-glass ceramics must be capable of combining with other products such as sintering ceramics, glazes, and glasses with special optical properties in a sintering process. With compositions such as 4866 SiO2 , 520 Al2 O3 , 315 K2 O, 3 20 Me(II)O (CaO, MgO, SrO, BaO), 0.55 P2 O5 , 312 Na2 O, it is possible to combine translucency with opalescence. The microstructure of this opal glass ceramic shows liquidliquid phase separation and leucite.
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics (d ) Twofold-nucleation mechanisms

583

Two possibilities for controlling double nucleation were discovered with the aim of precipitating two dierent crystalline phases in a glass ceramic by in situ reactions. First, a micaapatite glass ceramic was developed in the SiO2 Al2 O3 MgONa2 O K2 OP2 O5 F system (Vogel & H oland 1987). The base glass showed three glassy phases, two droplet phases and a glassy matrix. The small-droplet phase was rich in alkali uorine, and mica crystals formed in the glassy matrix by a solid-state reaction. The large-droplet phase was rich in CaO, P2 O5 and F and showed homogeneous nucleation of apatite. Therefore, during heat treatment of this glass between 750 and 1100 C, twofold nucleation was initiated: mica was nucleated heterogeneously and apatite homogeneously. Each apatite crystal grew within the droplet phase and growth was stopped at the boundary with the glassy matrix. The nal distribution of apatite was hexagonal crystals of diameter approximating that of the previous droplet phase. A second twofold nucleation mechanism was developed by combining surface and volume nucleation in the SiO2 Al2 O3 CaONa2 OK2 OP2 O5 F system. The result was the formation of a leuciteapatite glass ceramic (H oland et al . 1994). Apatite was precipitated as needle-like uorapatite. In recent publications, other needle-like apatites have been discussed in white opaque-glass ceramics, for example by Moisescu et al . (1999) or apatitemullite needles (Cliord & Hill 1996). A characteristic composition of leuciteapatite glass ceramic is 4958 SiO2 , 1119 Al2 O3 , 923 K2 O, 110 Na2 O, 212 CaO, 0.56 P2 O5 , 0.22.5 F, with additives of up to 6 wt% of CeO2 , B2 O3 , Li2 O. A CaO/P2 O5 molar ratio of ca. 5.8 and a uorine content of 0.6 wt% of the base glass composition were preferred. The controlled nucleation and crystallization of this base glass was carried out with glass powder of an average grain size of 2040 m. Nucleation of the glass and the sintering processes of the glass powder leading to a monolithic body proceeded almost simultaneously at temperatures between 800 and 1100 C. The twofold nucleation mechanism resulted in the formation of the two main crystal phases leucite and apatite. Leucite was produced by surface nucleation ( 3 c), while apatite was produced by volume nucleation combined with glass immiscibility. The mechanism of apatite formation in leucite-glass ceramics is dierent from that in the micaapatite glass ceramic. In the leuciteapatite glass ceramic, the stage of glass immiscibility is rapidly surpassed and the apatite quickly grows past the phase boundary of the amorphous droplet. The nucleation of apatite is heterogeneously initiated by crystalline precursor phases. The primary crystal phase is NaCaPO4 . The sequence of phase formation was studied in special monolithic glasses using high-temperature X-ray diractometry and nuclear magnetic resonance (Chan et al . 2001). Because of the reduced surface nucleation rate of leucite, it is easier to study nucleation of apatite in monolithic glasses than it is in glass powders. The composition of these bulk samples was 54.8 SiO2 , 14.1 Al2 O3 , 8.4 Na2 O, 10.6 K2 O, 4.9 CaO, 1.0 ZrO2 , 0.3 TiO2 , 3.9 P2 O5 , 0.8 CeO2 , 0.2 Li2 O, 0.3 B2 O3 and 0.7 F (H oland & Beall 2002). It was very surprising that the primary crystal phase of NaCaPO4 was formed in the base glass after quenching the melt. This crystal phase was determined by high-temperature X-ray diractometry up to ca. 610 C. The microstructure (gure 5) shows rounded crystals comparable in shape with phase separation of glassy droplet phases. Above 610 C, an additional crystalline phase
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

584

W. H oland and others

128.7 nm

98.4 nm

1 m
Figure 5. SEM of the etched (2.5% HF, 10 s) surface of primary crystals of NaCaPO4 in an apatiteleuciteglass ceramic after heat treatment for 15 min at 580 C. (Reproduced from Hoeland et al . (2002) with permission of The American Ceramic Society, copyright 2002. All rights reserved.)

2 m
Figure 6. Microstructure of a leuciteapatite glass ceramic for dental restorations. SEM of material heat treated for 1 h at 850 C and 1 h at 1050 C and then etched (3% HF, 10 s). (Reproduced from Hoeland et al . (2000c) with permission of Kluwer Academic Publishers.)

precipitated at 640 C and NaCaPO4 crystals were no longer present (table 2). These phases nucleate apatite heterogeneously. The rst crystals of apatite were identied after 8 h at 700 C. These crystals did not show a needle-like habit. Needle-like uorapatite was precipitated at 700 C and an additional heat treatment of 1050 C for 2 h. The growth of needle-like apatite was characterized as an Ostwald ripening process.
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics


Table 2. Crystal-phase formation in leuciteapatite glass ceramics temperature 25 C a 640 C 700 C 800 C

585

crystal phase NaCaPO4 Na2 Ca4 (PO4 )2 SiO4 KAlSi2 O6 , Na2 Ca4 (PO4 )2 SiO4 KAlSi2 O6 Na2 Ca4 (PO4 )2 SiO4 Ca5 (PO4 )3 F KAlSi2 O6 , needle-like Ca5 (PO4 )3 F

1050 C
a

After casting and cooling.

500 nm
Figure 7. SEM of the etched (3% HF, 10 s) surface of uorapatite glass ceramic, heat treated for 4 h at 520 C and 1 h at 700 C, showing the nanoscale microstructure.

In glass powders, the nucleation of leucite and apatite occurred in parallel during heat treatment for 1 h at 850 C. The leucite crystals were 12 m in diameter. The crystalline content of the glass ceramic was 1025 vol.% leucite and 510 vol.% apatite. The considerably smaller needle-like apatite crystals were located between the leucite crystals (gure 6). The morphology of these apatite crystals is comparable with that found in natural teeth. The glass ceramic is characterized by good optical properties, chemical durability and easy processing by sintering. It is applied as a biomaterial in restorative dentistry. (e ) Nature as the example for a glass-ceramic microstructure Apatite crystals of needle-like morphology are known to be contained in natural bone and teeth. The very small crystals in dental microstructures result in very special optical properties such as translucence and opalescence. Therefore, it was the aim
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

586

W. H oland and others

500 nm
Figure 8. SEM of glasses based on the SiO2 Li2 OK2 OZnOCaOP2 O5 F system permit crystallization of biomimetic structures: needle-like uorapatite after 4 h heat treatment at 700 C (etching 3% HF, 10 s).

of a research programme to develop a glass ceramic showing a similar microstructure to that of the natural material. This should provide some direction for the development of biomimetic microstructures. The main goal was to develop a needle-like, chemically durable apatite in a translucent glass ceramic that is easy to use as a dental restorative. A glass ceramic with needle-like uorapatite crystals precipitated in a glass matrix was developed in the SiO2 Li2 OK2 OZnOCaOP2 O5 F system (Schweiger et al . 2002). A typical composition range was 5665 SiO2 , 1.85.3 Li2 O, 917.5 K2 O, 9 16 ZnO, 3.510.5 CaO, 26 P2 O5 and 0.51.0 F. The glass of the given overall chemical composition with 6.1 CaO, 2.8 P2 O5 and 0.8 wt% F was heat treated for 4 h at 520 C and for 1 h at 700 C. The result was nanoscale (3060 nm) crystals of uorapatite Ca10 (PO4 )6 F2 uniformly and densely precipitated in the volume of the sample (gure 7). It was interesting to note that the microstructure formed by this heat treatment was not inuenced by variation in the content of CaO or P2 O5 . Higher concentrations of CaO (8.7 wt%) and P2 O5 (5.1 wt%) increased neither the size nor the volume percentage of the uorapatite phase. Based on the inuence of CaO, P2 O5 and F, it is obvious that phase separation provides the basis of the nucleation in the glass. However, primary crystal nucleation comparable with that in micaapatite glass ceramic (homogeneous nucleation of apatite in the droplet phase) or in leuciteapatite glass ceramic (heterogeneous nucleation by a precursor crystal phase) was not detected. It is most likely that glass immiscibility (glass-in-glass phase separation) and nucleation is so fast in this type of glass that it is not possible to distinguish the dierent reaction steps. The result is the formation of a glass ceramic microstructure with nanoscale uorapatite crystals, which do not show anisotropic needle-like growth in this temperature range. The process of nanophase formation was skipped, with a one-step heat treatment at 700 C. Crystal growth of needle-like uorapatite was observed by scanning elecPhil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics

587

tron microscopy. These results on monolithic glass ceramics demonstrate that, after short thermal treatments, the uorapatite crystals were smaller and more numerous than after long heat treatments at 700 C. One requirement for the crystallization process of Ostwald ripening is fullled with this result. The uorapatite crystals grew anisotropically in one direction to form elongated crystals (gure 8). The diameter of the elongated crystals remained constant at ca. 60 nm. The length increased from 120 to 300 nm, giving an aspect ratio of up to ve. The glass ceramic has good chemical durability and translucence. It is used as a highly aesthetic material to veneer metal-free dental restorations of the lithium disilicate type.

4. Conclusion
Fundamental and applied research activities have shown dierent possibilities for controlling nucleation in glasses to tailor properties. Volume nucleation enables the development of high-strength materials or nanophase products with zero thermal expansion in a wide temperature range. A twofold nucleation mechanism based on homogeneous nucleation in phase-separated glasses and on heterogeneous nucleation allows dierent crystal phases to be combined within a material. Other possibilities for twofold nucleation arise from the combination of surface and volume nucleation, resulting in a leuciteapatite glass ceramic for dental applications. The development of glass ceramics with needle-like apatite crystals demonstrates the applicability of biomimetic microstructures to a commercial glass ceramic.

References
Beall, G. H. 1992 Design and properties of glass-ceramics. A. Rev. Mater. Sci. 22, 91119. Beall, G. H. 1993 Glass-ceramics: recent development and application. In Nucleation and crystallization in glasses and liquids (ed. M. C. Weinberg). Ceramic Transactions Series, vol. 30, pp. 241266. Westerville, Ohio: The American Ceramic Society. Beall, G. H. & Duke, D. A. 1983 Glass-ceramic technology. In Glass science and technology (ed. D. R. Uhlmann & N. J. Kreidl), vol. 1, pp. 404445. Orlando: Academic. Beall, G. H., Karstetter, B. R. & Rittler, H. L. 1967 Crystallization and chemical strength of stued -quartz glass-ceramic. J. Am. Ceram. Soc. 50, 6774. Chan, J. C. C., Ohnesorge, R., Meise-Gresch, K., Eckert, H., H oland, W. & Rheinberger, V. 2001 Apatite crystallization in an aluminosilicate glass matrix: mechanistic studies by X-ray powder diraction, thermal analysis, and multinuclear solid-state NMR spectroscopy. Chem. Mater. 13, 41984206. Cliord, A. & Hill, R. 1996 Apatite-mullite glass-ceramic. J. Non-Cryst. Solids 196, 346351. Deubener, J., Br uckner, R. & Sternitzke, M. 1993 Induction time analysis of nucleation and crystal growth in di- and metasilicate glasses. J. Non-Cryst. Solids 163, 112. Gutzow, I. 1980 Induced crystallization of glass-forming systems: a case of transient heterogeneous nucleation. 1. Contemp. Phys. 21, 121137. Headley, T. J. & Loehman, R. E. 1984 Crystallization of a glass-ceramic by epitaxial growth. J. Am. Ceram. Soc. 67, 620625. H oland, W. & Beall, G. H. 2002 Glass-ceramic technology. Westerville, OH: The American Ceramic Society. H oland, W., Frank, M., Schweiger, M. & Rheinberger, V. 1994 Development of translucent glass-ceramics for dental application. Glastech. Ber. Glass Sci. Technol. C 67, 117122. H oland, W., Frank, M. & Rheinberger, V. 1995 Surface crystallization of leucite in glass. J. Non-Cryst. Solids 180, 292307.
Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

588

W. H oland and others

H oland, W., Frank, M., Schweiger, M., Wegner, S. & Rheinberger, V. 1996 Glass development and controlled crystallization in the SiO2 Li2 OZrO2 P2 O5 system. Glastech. Ber. Glass Sci. Technol. 69, 2533. H oland, W., Schweiger, M., Frank, M. & Rheinberger, V. 2000a A comparison of the microstructure and properties of the IPS Empress R 2 and the IPS Empress R glass-ceramic. J. Biomed. Mater. Res. 53, 297303. H oland, W., Schweiger, M., Cramer von Clausbruch, S. & Rheinberger, V. 2000b Complex nucleation and crystal growth mechanisms in applied multi-component glass-ceramics. Glastech. Ber. Glass Sci. Technol. C 73, 1219. H oland, W., Rheinberger, V., Wegner, S. & Frank, M. 2000c Needlelike apatiteleuciteglass ceramic as a base material for veneering of metal restorations in dentistry. J. Mater. Sci. Mater. Med. 11, 17. Holland, D., Iqbal, Y., James, P. & Lee, B. 1998 Early stages of lithium disilicate glasses containing P2 O5 : an NMR study. J. Non-Cryst. Solids 232234, 140146. James, P. F. 1982 Nucleation in glass forming systems a review. Advances in ceramics (ed. J. H. Simmons, D. R. Uhlmann & G. H. Beall), vol. 4, pp. 148. Columbus, OH: The American Ceramic Society. Kingery, W. D., Bowen, H. K. & Uhlmann, D. R. 1975 Introduction to ceramics. Wiley. Kokubo, T. 1991 Bioactive glass-ceramics properties and application. Biomaterials 12, 155163. McMillan, P. W. 1979 Glass-ceramics, 2nd edn. Academic. Moisescu, C., Carl, G. & R ussel, C. 1999 Glass-ceramics with dierent morphology of uorapatite crystals. Phosphorus Res. Bull. 10, 515520. Pannhorst, W. 2000 Surface nucleation. Charleroi, Belgium: International Commission on Glass. Petzoldt, J. 1967 Metastabile Mischkristalle mit Quarzstruktur im Oxidsystem Li2 OMgO ZnOAl2 O3 SiO2 . Glastech. Ber. 40, 385396. Schweiger, M., Frank, M., Cramer von Clausbruch, S., H oland, W. & Rheinberger, V. 1998 Microstructure and properties of pressed glass-ceramic core to zirconia post. Quintessence Dent. Technol. 21, 7379. Schweiger, M., Cramer von Clausbruch, S., H oland, W. & Rheinberger, V. 2000 Microstructure and mechanic properties of lithium disilicate glass-ceramic in the SiO2 LiO2 K2 OZnOP2 O5 system. Glastech. Ber. Glass Sci. Technol. 73, 4350. Schweiger, M., H oland, W. & Rheinberger, V. 2002 Nanophase formation in dierent glassceramic systems. In Proc. American Ceramic Soc. Meeting 2002, St Louis, MO. Ceram. Transactions 137 American Ceram. Soc. Stookey, S. D. 1959 Catalyzed crystallization of glass in theory and practice. Ind. Engng Chem. Res. 51, 805808. Uhlmann, D. R. 1980 On the internal nucleation of melting. J. Non-Cryst. Solids 41, 347357. Uhlmann, D. R. & Kolbeck, A. G. 1976 Phase separation and the revolution in concepts of glass structure. Phys. Chem. Glasses 17, 146158. Vogel, W. & H oland, W. 1987 The development of bioglass ceramics for medical application. Angew. Chem. Int. Ed. Engl. 26, 527544. Weinberg, M. C., Bernie III, D. P. & Shneidman, V. A. 1997 Crystallization kinetics and the JMAK equation. J. Non-Cryst. Solids 219, 8999. Zanotto, E. D. 1994 Crystallization of glass: a ten year perspective. 1993 Vittorio Gottardi Prize Lecture. Chim. Chron. New Ser. 23, 317.

Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

Control of nucleation in glass ceramics Discussion

589

K. F. Kelton (Department of Physics, Washington University, St. Louis, MO, USA). Are elemental metals still used for nucleants? If not, why not? land. Metals such as Au, Ag or Pt were used in the 1960s and 1970s as nucleW. Ho ating agents to precipitate crystals in base glasses to form glass ceramics. Today, this type of heterogeneous nucleation is used very rarely. Now we are able to control nucleation and crystallization of glasses by surface or volume mechanisms. It is very important to combine heterogeneous volume nucleation with phase separation processes. A special micro-immiscibility is possible with the following nucleating agents: TiO2 , ZrO2 , P2 O5 , Ta2 O5 , WO3 , Fe2 O3 and F. These agents allow the formation of uniform crystals within the glassy matrix and to control the crystal content of the glass ceramic. Therefore, it is possible to produce products with tailor-made properties. B. R. Haywood (Department of Physics, Keele University, UK ). Given that the ux electrolytes are critical, have you considered the potential role of zeolites as reagents to regulate the crystallization process? land. There is the possibility of precipitating zeolite-type crystals in glass W. Ho ceramics. In forming these crystals or any other type of crystals, the viscosity of the glassy matrix changes and the following crystallization is inuenced. But there is not now any special inuence of zeolites in controlling crystallization of glasses at high temperatures.

Phil. Trans. R. Soc. Lond. A (2003)

Downloaded from rsta.royalsocietypublishing.org on September 10, 2012

You might also like