Sunteți pe pagina 1din 33

1.

01 Polyacrylonitrile (PAN)-based Carbon Fibers


A. SHINDO Fiber Science Laboratory, Hyogo, Japan
1.01.1 INTRODUCTION 1.01.2 PROCESSING 1.01.2.1 Precursors for CFs 1.01.2.2 Stabilization 1.01.2.3 Carbonization and High-temperature Heat Treatment 1.01.2.4 Surface Treatment 1.01.2.4.1 Anodic oxidation 1.01.2.4.2 Plasma treatment 1.01.2.5 Ceramic Coatings 1.01.3 STRUCTURE 1.01.3.1 High-strength CFs 1.01.3.2 High-modulus CFs 1.01.3.2.1 Longitudinal sections 1.01.3.3 Structural Parameters and Density 1.01.3.3.1 Morphology of fracture surface 1.01.3.4 Radial Heterogeneity 1.01.3.5 Schematic Structure 1.01.3.6 Chemical Composition 1.01.4 MECHANICAL PROPERTIES 1.01.4.1 Longitudinal 1.01.4.1.1 Elasticity 1.01.4.1.2 Tensile modulus 1.01.4.1.3 Tensile strength 1.01.4.1.4 Compressive strength 1.01.4.2 Transverse 1.01.4.2.1 Transverse modulus 1.01.4.2.2 Torsional modulus 1.01.4.3 High-temperature Properties 1.01.5 ELECTRIC AND MAGNETIC PROPERTIES 1.01.5.1 Electrical Resistance and Thermoelectric Power 1.01.5.2 Electromechanical Properties 1.01.5.3 Magnetoresistance 1.01.6 THERMAL PROPERTIES 1.01.6.1 Thermal Expansion 1.01.6.2 Thermal Conductivity 1.01.7 SURFACE PROPERTIES 1.01.7.1 Morphology and Surface Areas 1.01.7.2 Functional Groups 1.01.7.3 Surface Free Energy 2 2 2 3 5 7 7 8 8 9 10 10 10 10 11 12 13 14 14 14 14 15 16 18 20 20 21 22 23 23 24 24 24 24 25 26 26 26 28

2
1.01.7.4 Wetting Property 1.01.7.5 Reactivity 1.01.8 REFERENCES

Polyacrylonitrile (PAN)-based Carbon Fibers


28 30 31

1.01.1

INTRODUCTION

1.01.2

PROCESSING

PAN-based carbon fiber (CF) occupies a premier position among high-performance fibers for composites. This was released into the market in the 1960s (Shindo, 1961). The first decade of its use may be regarded as a period of incubation. During the second and third decades, the CF has experienced remarkable growth in producing technology, particularly with respect to tensile strength. It is considered that the factors forming the basis of the technology in the early stage are: (i) the inherent nature of PAN fiber, particularly that which exhibits preferred orientation of molecular fragments or basic structural units at every stage of heat treatment for carbonization (Shindo, 1961, 1964); (ii) the air-oxidizing stabilization treatment (Shindo, 1961, 1964); (iii) suppression of length shrinkage of fibers or stretching for increasing or keeping the preferred orientation of molecular fragments or graphitic layer planes; (iv) elimination of flaws (Johnson, 1969; Sharp and Burney, 1971; Moreton and Watt, 1974; Reynolds and Moreton, 1980); and (v) anodically oxidizing surface treatment for use as a reinforcement (Courtaulds Ltd., 1964). Watt and Johnson (1969), who started their study in autumn 1963 (Anon., 1970), pointed out that stretching fibers during the stabilization step is effective for increasing the Young's modulus and tensile strength of CFs. The growth of the CF industry was primarily driven by the development of its application in aerospace and sports industries. Especially, high-performance requirement for CFs in the aerospace industry led CF manufacturers to the development of CFs with markedly higher fracture strain during the latter half of the 1980s. In response to the development of higher performance CFs and the expectation for further development of their application, there has been also an advance in scientific studies on that structure and mechanical properties. In this chapter, such processing, structure, and various properties will be reviewed. As regards references on CFs, there are books written by Fitzer (1985), Watt and Perov (1985), Dresselhaus et al. (1988), Donnet and Bansal (1990), and Peebles (1994).

PAN-based CFs are manufactured by making tensioned endless yarns or tows of acrylic fibers passed continuously through a series of zones for stabilization, carbonization, and surface treatment. For making HM CFs, the fibers are subjected to high-temperature heat treatment just after the carbonization treatment. A final treatment temperature is selected according to desired properties such as tensile strength and Young's modulus. Thus, high-strength, intermediate-modulus, or high-modulus type CFs are produced. The surface-treated CFs are usually sized with epoxy resin.

1.01.2.1

Precursors for CFs

PAN fibers used as precursors for CFs are made of copolymers containing acrylonitrile in excess of 90%. Pure PAN has a melting point of 326 8C (Dunn and Ennis 1970) or a slightly lower temperature (Layden, 1970), while copolymer PAN usually has melting points lower than that. Such a high melting point makes the stabilizing treatment of PAN fibers possible. Acrylic fibers of copolymer with comonomerbearing carboxylic groups such as itaconic acid (Grassie and McGuchan, 1972)

COOH CH2 C CH2 COOH


or hydroxylic groups such as hydroxymethylacrylic compound (Morita et al., 1973, 1972; Takahagi et al., 1986)

H2C=C R

CHOH

wherein X stands for a radical selected from a certain group, and R is a member of another certain group, are recommended for use as precursor fibers. It is assumed that these comonomers are incorporated to initiate nitrile

Processing

Figure 1 Gravimetric analysis curves in nitrogen gas showing an effect of preoxidation with air at 200 8C by increasing the yield of CF. *: preoxidized, *: not preoxidized (after Shindo, 1961).

usually performed at temperatures between 200 (Shindo, 1961) and 350 8C (Mittal et al., 1997a), in air flows for about 30 min or more. Even in air, if acrylic fibers are rapidly heated, drastic exotherm occurs, which results in violent degradation of the fiber. To avoid these phenomena, the heating in air must be started at a temperature below that at which the exotherm begins, i.e., around 200 8C. Incorporation of the comonomers mentioned above into the polymer is also very effective for checking violent exotherm reactions (Grassie and McGuchan, 1972). In addition, it is desirable that the fibers are heated at increasing temperatures at controlled rates. During the stabilization step, polymerization of nitrile groups and dehydrogenation from the main chains and addition of oxygen to them occur to form a stabilized structure in the fiber. As regards nitrile polymerization or cyclization, a number of reactions have been proposed, including both intra- and interchain reactions (Bashir, 1991; Peebles, 1994). Intrachain cyclization into condensed rings comprising a few rings, however, seems to have become the accepted mechanism:

H2 C C CN C CN

H2 C C CN H2 C

H2 C C CN H2 C C C N N H2 C C C N

polymerization (Grassie and McGuchan, 1972), very helpful to stabilization. As regards the thickness of fiber, the thinner the precursor fiber, the higher the preferred orientation of molecules along the fiber axis generally becomes. The high molecular orientation in precursor fiber generally brings about high tensile modulus and tensile strength of CFs. Precursor fibers, therefore, are extensively stretched during the spinning operation

C C

C C

1.01.2.2

Stabilization

The (oxidative) stabilization is indispensable for converting PAN precursor fibers into fibers that can withstand the rigors of the carbonization processing and a crucial step in the manufacture of PAN-based CFs. One of the most striking effects of the stabilization treatment is an increase in carbon yield, as shown in Figure 1 (Shindo, 1961). The stabilization treatment is

Furthermore, it remains an unsolved problem whether dehydrogenation from the main chains of PAN precedes or follows the nitrile polymerization, or occurs at the same time. Nitrile polymerization is not only for stabilization, but dehydrogenation or addition of oxygen is essential for substantial stabilization. Some structures for the stabilized fiber have been proposed (Usami et al., 1990). One of them is the following (Takahagi et al., 1986):

OH

N H

N 40% 30%

20%

CN 10%

Polyacrylonitrile (PAN)-based Carbon Fibers

Figure 3 Percentage shrinkage at various loads along with DSC behavior on heating in air (18C min71) (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996, 34, 14271445).

Figure 2 (a) Irregular helical structure of PAN chains. (b) Morphological model of PAN fiber showing ordered and disordered regions (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996, 34, 14271445).

On heating in air, untensioned copolymer fibers which have been stretched during spinning shrink to a considerable extent. The fiber is held, therefore, under tension as high as the fiber strength allows to check the shrinkage in order to preserve the orientation. In this case, shrinkage stress develops in the fiber, exhibiting a peak at a temperature between 100 and 200 8C, and another peak between 200 and 300 8C (Warner et al., 1979). According to Gupta and Harrison (1997), intramolecular cyclization initiates in the amorphous parts of the fiber at around 200 8C, and develops at higher temperatures, randomization of the highly oriented amorphous phase taking place with progression of the stabilization. Figure 2 shows the structure of PAN chains and a morphological model of the PAN fiber (Warner et al., 1979).

Intramolecular cyclization and randomization in the amorphous phase spread to the crystalline regions at higher temperatures. The second rise of stress is due primarily to an entropic recovering force generated by randomization of the highly oriented amorphous and crystalline phases in which stabilization modifications of polymer molecules progress. In particular, randomization of the amorphous phase primarily contributes to the stress development, although randomization of crystalline phase may also contributes. In the stabilization process, some sort of molecular chain scission is unavoidable. On the other hand, at least at the stage at which stabilization has proceeded to some extent, some sort of intermolecular cross-linking takes place between cyclized segments, in particular in a fiber of a copolymer where the comonomer exerts a strong cyclization-initiating effect as mentioned previously. This is considered to have the lateral cohesive force of the fiber (Shindo, 1971). The stabilization reactions are accompanied by an exotherm. Figure 3 shows a differential scanning calorimetry curve of a fiber of an acrylic copolymer, methylacrylate, and itaconic acid on heating in air (Gupta and Harrison, 1996). At temperatures above 300 8C, it has been considered that the condensed rings, partially dehydrated or oxidized, cross-link to each other to form aromatic condensed rings. Thus, stabilization reactions proceed to completion at around 380 8C (Mittal et al., 1994; Gupta and Harrison, 1996). The stabilization reactions can also proceed to an almost full extent at temperatures between around 230 and 290 8C. Shrinkage stress when the fiber is held without length change (Hiramatsu et al., 1972; Warner et al., 1979; Takaku et al., 1981) can be used as

Processing

Figure 4 Tensile strength of CFs heat-treated to 1000 8C in nitrogen for 20 min under a tension of 1 MPa or 10 MPa as a function of SI values of the fibers oxidized in air under a constant length. The oxidation temperatures were 242, 253, or 264 8C (reproduced by permission of Elsevier Science Ltd. from Carbon, 1995, 33, 783788).

Figure 5 Effects of preoxidation and carbonizing atmosphere of HCl gas on numbers of atoms, per monomer unit of PAN, of the composing elements remaining in a PAN fiber. - - - - - : preoxidized, and carbonized in HCl gas, : preoxidized, and carbonized in nitrogen. &: unpreoxidized, and carbonized in nitrogen (after Shindo, 1971).

an in situ measure of the extent of stabilization, although the oxygen content in stabilized fibers has often been used as a measure. When a PAN fiber is cooled down from a temperature on the way to stabilization, the shrinkage stress of the fiber decreases, reaching a constant value at 180 8C. From this observation, Ogawa and Saito (1995) proposed a parameter (SI value) for evaluating the degree of oxidative stabilization of a PAN fiber from shrinkage stresses. Figure 4 shows the content of oxygen bound chemically in the stabilized fibers as a function of that parameter. A good linear relation is seen between them. In Figure 4, an SI value of 100% corresponds to a chemically bound oxygen of 12 wt.%, which has been proposed as an optimum value of oxygen to be contained in stabilized fibers for high-performance carbon fibers. Thus, shrinkage stresses are applicable for evaluating the degree of stabilization of an operation for the continuous manufacture of CFs. Stabilization of PAN fibers requires the longest processing times among the heating steps in a line for carbon fiber production. Several approaches to stabilization, therefore, have been researched. Some of them are methods based on making thermostable PAN fibers by treating with atmospheres containing nitrogen oxide (Morita et al., 1972), sulfur dioxide (Morita et al., 1972), active sulfur derived by photolysis of carbonyl sulfide, oxygen, and hydrochloric acid gas (Shindo, 1971), and ammonia and air (Bhat, 1990). Other approaches are pretreatment with potassium permanganate (Ko et al., 1988; Mathor, 1994), acetic acid or succinic acid solution (Mittal et al., 1997b), solutions

of amines, basic polyol solutions and cuprous chloride solutions (Peebles, 1994), prior to air oxidation.

1.01.2.3

Carbonization and High-temperature Heat Treatment

Stabilized fibers are heated in order to be carbonized at temperatures in the range 1000 1600 8C in a nitrogen atmosphere under slight tension. During the carbonization process, carbon atoms or chains in the fibers condense into carbon material, increasing the carbon content in the fiber progressively as seen in Figure 5 (Shindo, 1971), while the other elements such as hydrogen, nitrogen, and oxygen are eliminated through the evolution of gases (Shionoya et al., 1972; Watt, 1972). From gas evolution ratetemperature profiles, the carbonization process can be divided into two regions, below and above about 700 8C, although this border temperature more or less varies depending on the degree of stabilization of the fiber and on the heating condition. In the region below 700 8C, decomposition of stabilized PAN molecular fragments occurs, and gases such as HCN, H2O, CO2, CO and NH3 evolve through primarily intramolecular reactions between functional groups. Intermolecular cross-linking between the fragments also proceeds to form condensed benzene aromatic rings through dehydration (Watt, 1970) and elimination reactions of hydrogen cyanide (Shionoya et al., 1972). This cor-

Polyacrylonitrile (PAN)-based Carbon Fibers As the heating temperature rises, those fragments increase in extent, bonding between the fragments becomes increasingly firm through such reactions, and the carbon atom network in the fiber becomes more and more dense, while the volume of open pores formed during the course of gas evolution decreases from around 1000 8C up to around 1200 8C of HTT (Spencer et al., 1970). At temperatures above 1200 8C, closed pores remain in the fibers. The tensile strength and Young's modulus of the fibers keep increasing up to an HTT of around 1500 8C or 1600 8C as the carbon atom condensation progresses, as shown in Figure 6 (Mittal et al., 1997a). Although the carbonization treatment is usually carried out in an inert atmosphere, it has been observed that HCl gas mixed in the atmosphere below 600 8C causes a considerable increase in carbon yield and tensile strength of the CFs obtained (Shindo, 1971). For CFs to be high modulus, we have to subject the fibers to a high temperature, such as 2500 8C or 2800 8C. During treatment at higher temperatures, aromatic layer planes in CFs increase in extent primarily through the coalescence of neighboring layers. The increase in the extent of layer planes brings about an

Figure 6 Tensile strength and Young's modulus vs. HTT curves (reproduced by permission of Elsevier Science Ltd. from Carbon, 1997, 35, 1196 1197).

responds to the phenomena that Young's modulus and tensile strength start to increase at around 300 8C (Shindo, 1961; Watt, 1970) or 400 8C (Mittal et al., 1997a). In the region above 700 8C, the following intermolecular reactions are supposed to occur, resulting in the formation of polynuclear aromatic molecular fragments through the evolution of N2 (Watt, 1970) and HCN gases (Shionoya et al., 1972):

N
N

N N

N N

N N

+ nN2

N N

N N N N

nHCN

Processing

Figure 7 Surface of Besfight HT anodically oxidized in sulfuric acid solution. (a) Washed with water. (b) Washed with alkaline solution and then with water (after Shindo, 1983).

increase of preferred orientation along the fiber axis. Those increases cause an increase in the Young's modulus of CFs. Thus, the Young's modulus keeps increasing with the rise in HTT, whereas the tensile strength decreases (Mittal et al., 1997a).

1.01.2.4 1.01.2.4.1

Surface Treatment Anodic oxidation

Almost all PAN-based CFs produced are used as reinforcements for resin matrix composite materials. These have therefore been treated by anodic oxidation to improve their adhesion with the resin matrix, following carbonization or high-temperature treatment

because anodic oxidation has advantages over other oxidation techniques in rate, uniformity, and controllability of the degree of oxidation. Although a number of electrolytes could be used for the anodic oxidation, alkaline electrolytes, for example, sodium hydroxide and ammonium bicarbonate, are recommended to be used because the degradation products formed on the fiber surface dissolve in an alkaline aqueous solution leaving no residues, which simplifies the fiber washing, unlike in acidic solution (Shindo, 1983). In acidic solution, the degradation products remain as residues on the fiber surface, without dispersing or dissolving into the solution. Figure 7 shows the surface of a fiber immediately after an anodic oxidation at 3450 C g71 in a sulfuric acid solution and washing with water, and the surface of the fiber washed with an alkaline aqueous solution and water. It can be seen in Figure 7(b) that the degradation products have been washed off from the fiber surface. The degradation products are considered to be polynuclear aromatic molecules with carboxylic and hydroxyl groups on their peripheries. In this case, the amount of degradation products was about 15%, about 70% of the total weight loss of the fiber. Ehrburger and Donnet (1985) have described that, when an HT-type CF is anodically oxidized in an alkaline electrolyte, carbon dioxide is formed together with degradation products which are dissolved in the electrolyte. Darkening of alkaline electrolyte solution during the anodic oxidation of CFs has been also observed by Kozlowski and Sherwood (1985). Furthermore, Wu et al. (1995) have removed the degradation residue on the surface of nitric acid-oxidized fibers by washing the fibers with aqueous sodium hydroxide. Furthermore, epoxy matrix composites of the HT and HM CFs oxidized in a continuous treatment line showed that the ILSS for the fibers oxidized in a sodium hydroxide solution was higher than for those in the acidic electrolyte. The lowering of the interlaminar shear strength (ILSS) of fiber reinforced plastics seems to be due to the degradation product residing on the fiber surface in the acidic solution (Shindo, 1983). Rises in ILSS with cell current and surface atomic concentrations of carboxylic acid and hydroxyl groups for CFs subjected to anodic oxidation in alkaline solution have been also reported by Harvey et al. (1987) and Alexander and Jones (1995), respectively. With commercial CFs, anodic oxidation increases the surface area slightly and the active surface area markedly, in addition to forming surface active groups such as carboxylic acid

Polyacrylonitrile (PAN)-based Carbon Fibers very close to that needed for anodic oxidation treatment, and for being usable to introduce amine groups onto CF surfaces. A low-power air plasma can introduce alcohol and carboxyl groups onto the surface of a HT CF, Amoco T300, and carboxyl groups onto the surface of a HM CF, Hercules HMU, whereas nitrogen and ammonia plasmas can form aromatic amines (NH2), andC=NH groups, which have the potential to bond with epoxy resins (Jones and Sammann, 1990a). It has been pointed out that those chemical changes mainly occur at the edge site or defects and not on the basal planes (Jones and Sammann, 1990b). Furthermore, plasma treatment can coat CFs with polymer. For example, treatment with allylcyanide or xylene/air/argon plasma has brought about an increase in the tensile strength of HT fibers, Hercules and Grafil. This increase is probably due to the filling of cracks and flaws by the plasma-deposited polymer on the fiber surface (Dilsiz et al., 1995). Amino groups react readily with carboxyl groups on the CF surface. With oxidized CFs, amino groups can also be introduced by reaction with carboxyl groups on the CFs (Pittman et al., 1997). Such a treatment is effective for enhancing adhesion of CFs to polyurethanes and epoxy resin matrices.

and hydroxide. As seen in Table 13, commercially used anodic oxidation increases the specific surface area by 34% for HS fibers and by about 20% for a HM fiber, and active sites by 7090% for HS fibers and by 240% for a HM fiber, whereas it scarcely decreases the tensile strength (Shindo, 1983). Harvey et al. (1987) have suggested that carboxylester groups are formed at edge sites in a CF surface with anodic oxidation, whereas ketoenol groups are formed on the basal planes. Nakahara and Shimizu (1992) oxidized anodically pyrolytic graphite specimens in an alkaline electrolyte solution, tetraethylammonium hydroxide, and found that on the edge surface there was no increase in the -COOH/C ratio, whereas the -OH/C ratio increased gradually with increasing electric charge up to 5000 C cm72, and oxidation was limited only to the edge surface. The anodic oxidation of CFs in an ammonium bicarbonate electrolyte, furthermore, causes formation of CN groups in CF surfaces (Proctor and Sherwood, 1983; Bradley et al., 1994). In this case, increasing the concentration of ammonium ions by saturating the solution with ammonia introduces larger amounts of surface nitrogen (Kozlowski and Sherwood, 1986). Alexander and Jones found that the formation of amide groups (CONH2), in addition to carboxyl, hydroxyl, and aromatic imide groups, on the surface of AU fiber, Hercules, during an anodic oxidation in ammonium bicarbonate electrolyte, and have also supposed that the amide groups result from reaction of the ammonium ion in the electrolyte with the surface acidic functionalities (Alexander and Jones, 1996). With low-modulus CFs, Courtaulds LM and AKZO HTA, amide and/ or aromatic amines (PhNH2) seem to be formed during anodic oxidation. The oxidation of Courtaulds high-modulus CFs also appears to lead to the formation of aliphatic amine and protonated amine (Bradley et al., 1994), in addition to amide groups. It is further possible to introduce amino groups onto CF surface by reaction of caboxylic acid groups in the surface with tetraethylenepentamine (Pittman et al., 1997). 1.01.2.4.2 Plasma treatment

1.01.2.5

Ceramic Coatings

In addition to anodic oxidation, there are many methods available to alter the surface chemistry of CFs in an attempt to promote chemical bonding between the fiber and resin in composite materials. Among them, plasma treatment is noticeable from the point of view of treatment time as short as 1530 s, which is

CFs can also be used as reinforcements for inorganic matrices (for a review see Peebles, 1994). The following two cases of composites particularly become an issue. First are the composites with brittle matrices having a lower fracture strain than those of the CFs, and second are those with matrices which form a brittle carbide layer through reactions with CFs on fabrication. In the former, cracks formed in the matrix under tension will propagate into the fibers if the interfacial adhesion between fiber and matrix is strong enough to withstand the shear stress at the interface. For a CF with a brittle layer of a thickness above a certain value, a similar mechanism acts in crack propagation from the layer into the fiber. CFs, when heated in a flow of a gaseous mixture of TiCl4, hydrogen, and argon at a temperature from 800 to 1000 8C, are coated with titanium carbide (Honjo and Shindo, 1986a). Adhesion of the TiC layer formed in this way with the core CF is naturally strong. The relation of the strength of the TiC-coated fiber and the layer thickness accords to the equation, sm ! d71/2. This equation comes from the equation, sm = k(2Eg/pd)1/2, where k is a factor related

Structure

Figure 8 Variation of tensile strength of double layer coated CFs by interfacial debonding stress (reproduced by permission of Elsevier Science Publishing Co., Inc. from `Proceedings of the 1st International Conference on Composite Interfaces', 1986, pp. 101107).

It is known from these facts that the adhesive strength between the carbon layer and ceramic layer is lower than that between the carbon layer and CF. In fact, the end of the doubly coated fiber pulled out from resin exhibited an electron diffraction pattern characteristic of the carbon coating (Honjo and Shindo, 1986b). Murakami et al. (1986) have also observed, by using an X-ray energy dispersive analyzer, that when the aluminum wire reinforced with carbon + SiC or carbon + SiC + TiB layercoated fiber fractures under tension, a peeling off occurs at an interface either between the carbon and SiC layers, or between the carbon layer and SiC + TiB layer. These results suggest that the surface of carbon coating is less active than that of the bare fiber. In fact, the ASA of the fibers coated with a carbon layer alone was only about one-fifth to one-third that of the bare fiber (Shindo and Honjo, 1986).

to the shape of the crack (Ochiai and Murakami, 1979), E and g are the modulus of elasticity and the surface energy of the CF, respectively, and d is the thickness of the layer. The strength of the fiber coated with films with thicknesses above 10 nm decreases with increasing thickness of the TiC layer according to the above equation. The strength of the fiber, however, regains its original strength when the TiC layer is removed from the fiber with a solution mixture of HNO3 and HF. CFs can also be coated with SiC. A SiC layer with a uniform thickness, about 0.5 mm, was formed by heating CFs in a flow of a gas mixture of monomethyltrichlorosilane, methane, hydrogen, and argon, at a certain hydrogen concentration, at a temperature of about 1200 8C (Honjo and Shindo, 1986b; Shindo and Honjo, 1986). With Torayca T300 fiber, the SiC coating resulted in a tensile strength loss of 40%. In contrast to this, the strength of the fiber coated with carbon layer was higher by about 20% than that of the bare fiber. These results led Honjo and Shindo to develop the technique of double and triple coatings. Thus, a Torayca T300 fiber yarn was continuously coated with a carbon layer and then with a ceramic layer such as TiC(1), TiC(2), SiC, or TiN. Figure 8 shows the variation of tensile strength of the CF coated with two layers with increasing interfacial debonding strength between the coating and fiber (Honjo and Shindo, 1986b). Further, the double layer coated fiber exhibited a lower debonding strength than that of the fiber coated with the corresponding ceramic layer alone.

1.01.3

STRUCTURE

CFs heat-treated in the 10002500 8C range are composed of turbostratic stacks of aromatic (graphitic) layers (basic structural units), shown in Figure 9, and disordered regions including cross-linking. The sizes of the layers or stacks, and the orientation along the fiber axis, increase as HTT rises. The aromatic layers connected longitudinally with cross-linking and/or disordered regions are wrinkled at HTTs in the

Figure 9 (a) Crystal structure of graphite crystal. (b) Structure of turbostratic carbon.

10

Polyacrylonitrile (PAN)-based Carbon Fibers 1.01.3.1 High-strength CFs

TEM studies have showed that HT CFs are made of very small BSUs (about 10 A) (Oberlin and Guigon, 1988; Dobb et al., 1995), as shown in Figure 10. The studies performed by all possible modes of TEM show that highstrength type CFs have very small (about 100 A) local molecular orientations of BSU (Oberlin and Guigon, 1988). According to them BSU, in transverse sections, is less than 10 A in diameter and thickness. The direction of the fringes changes continuously. An intricate and entangled structure is thus formed. On the basis of these observations, in the model of microstructure proposed by Guigon et al. (1984) for a HT fiber, the BSU are associated in a zigzag form with tilt and twist boundaries. Thus, they form larger wrinkled sheets. When two randomly crumpled sheets of BSU come close enough together, either transversally or longitudinally, bonding occurs in the faulty areas at the places where they touch each other. Cross-linking atoms, tetrahedral bonds, etc. that are frozen in at the BSU boundaries are responsible for the fiber cohesion. Among high-strength CFs, Lc (2025 A) or Z (SAD) (3540 A) do not vary significantly (Oberlin and Guigon, 1988).

1.01.3.2 1.01.3.2.1
Figure 10 (a) 002 Dark-field of a longitudinal section of Toray 300, 40B 2600 221. (b) Schematic representation of BUS preferred orientation. (c) 002 Lattice fringe image of Serofim AXT 2666, to be compared to (b) (reproduced by permission of Elsevier Applied Science Publishers Ltd. from Fibre Science and Technology, 1984, 20, 177198).

High-modulus CFs Longitudinal sections

15002000 8C range, and are planar at HTTs above 2000 8C (Oberlin and Guigon, 1988.). Most PAN-based CFs are unable to graphitize. Nevertheless, the regions of large layer planes, where the cross-linking with neighbors is weak, can be slightly graphitized. In fact, the arcs (112), (103), and (114), which correspond to the planes of the three-dimensional lattice in the graphite crystal, have appeared in the electron diffraction photographs of CFs heat-treated at 2800 8C and 3000 8C, though they are faint, in addition to the arc (101) (Shindo, 1961a, 1961b). Further, there are microfibrils in CFs (Johnson and Watt, 1967; Kwizera et al., 1982). The dimension of microfibrils reaches about 100 nm or more according to types of CFs.

HM CFs are made of isometric distorted lamellae (Oberlin and Guigon, 1988). The lamellae measured in 11 dark-field images have always been found to be isometric (La\ = Lak). 002 lattice fringes for HM CFs show stacks of quasi perfect fringes (Oberlin and Guigon 1988). There are also reviews by Jain and Abhiraman (1987) and by Johnson (1987) on the structure of CFs.

1.01.3.3

Structural Parameters and Density

In general, the structural parameters, such as degree of preferred orientation along the fiber axis and crystallite sizes, increase, and the d-spacing (002) decreases with increasing HTT. The fiber density increases as the HTT rises. Such structural parameters for types of CFs are shown in Table 1 (Kumar et al., 1993; Dobb et al. 1995). The densities of commercial CFs are shown in Table 2. According to Oberlin and Guigon (1988), whatever the fiber examined, Lc002 reaches 30 A, and LaTEM (La by TEM) at an HTT of 1700 8C. The latter 150200 A

Structure
Table 1a Fiber T-300 T-40 T-50 GY-70 AS-4 M40J M60J IM8 Z (8) 35.1 30.2 16.4 9.6 36.8 21.4 9.9 d (002) 0.342 0.343 0.3423 0.3396 0.342 0.3427 0.3411 0.3431 Structural parameters of various CFs. Lc (nm) 1.5 1.8 5.3 14.1 1.8 3.6 7.8 1.9 La (0) 2.2 3.0 6.5 12.0 3.0 5.5 7.0 3.1 La (90) 4.1 4.6 17.5 40.0 4.0 12.5 28.0 5.1 Three-dimensional order No No Maybe Yes No No Yes No SEM morphology Particles Particles Some sheet-like Sheet-like Particles Particles

11

Source: Kumar et al. (1993). Z: Full-width at half-maximum of the (002) azimuthal scan, d(002): d-spacing of crystallites, Lc: Crystallite size perpendicular to graphitic basal plane, La(0) and La(90): Crystallite sizes along graphitic plane perpendicular and parallel to the fiber axis.

Table 1b Fiber C1 C2 X Y Z T1000 Z(8) 20.1 36.1 21.2 22.0 30.8 31.5

Structural parameters of various CFs. Dd 55.8 66.0 50.2 49.9 57.5 62.2 Dc 33.3 86.7 40.0 40.0 80.0 86.7 La|| 4.6 3.0 4.8 4.8 3.4 2.9 La\ 7.4 5.3 7.2 7.0 5.9 5.2 Lc 5.9 1.7 4.7 4.3 2.6 1.7

Source: Dobb et al. (1995). Dd: Intercrystallite disorder, Dc: Intracrystallite disorder, La||: Crystallite length along the a-axis, La\: Crystallite width perpendicular to the a-axis.

increases from 200250 A at 2000 8C to 250 300 A at 2200 8C, and then reaches about 400 A at 2800 8C, whereas Lc002 remains near to 35 A (Oberlin and Guigon, 1988). 1.01.3.3.1 Morphology of fracture surface

According to Vezie and Adams (1990), there is a distinct difference in the morphology of a fracture surface between lower modulus and high modulus PAN-based fibers when observed by a high-resolution SEM. Lower modulus fibers, AS-4. T-300 and T-40, G40-700, G45-700, and IM-8, show a fracture surface exhibiting rough, rather poorly defined granular textures, with no indication of sheet-like or fibrillar structures, whereas a HM fiber, GY-70, is made up of sheet-like structures. A transition structure of very fine sheets is seen in an intermediate modulus T-50 fiber. A low magnification image of this fiber demonstrates the presence of core and sheath regions, except that in this case, the rough, granular texture of the core and sheath, typical of HT PANbased fibers, is separated by fine, radially aligned sheets. T-50 fiber with such a transi-

tion structure between the sheet-like structure and the granular texture has intermediate mechanical properties. Kogure et al. (1994) also observed a fine granular microstructure on fracture crosssectional surfaces of Hercules HM3000 and postcreep (plastically deformed by 16.5% at 2310 8C). Some sheet-like characteristics were also observed to develop in T50 fiber (Kumar et al., 1993). High-resolution SEM micrographs of the transverse sections of T-40 and M-60J fibers also do not show the development of sheet-like morphology, but exhibit a particulate morphology. The morphology of T-40 fiber appears to be quite inhomogeneous. Careful examination of a SEM photograph of this fiber reveals the existence of some very small particles of about 20 nm in size. The particle size in M 60J is approximately 4050 nm, and this fiber also exhibits the absence of sheet-like character, though the modulus is higher than that of GY-70. Both fibers, however, show three-dimensional order. Good agreement is observed between the sheet thickness and crystallite size, Lc, in GY-70 fiber, both having values of about 15 nm (Kumar et al., 1993).

12
Table 2 Mfr.

Polyacrylonitrile (PAN)-based Carbon Fibers


Mechanical and other properties of types of CFs extracted from manufacturers' data sheets. Fiber type T-300 T-40 T650/42 T-50 AS4 IM4 IM7 UHM TR30 TR50 MR50k SR50 C30 C35 HTA UTS IMS UMS HTA ST4 IM600 HM35 TM40 UM68 T300* T300J* T700S T800H* T1000G M40J* M50J M60J X665 M40* Panex33 Panex30 Filament count 1k, 12k 12k 6k, 12k 3k, 6k 3k, 12k 12k 6k, 12k 3k, 12k 3k 12k 12k 12k 1k, 24k 12k 6k, 24k 12k 3k, 12k 12K 12k, 24k 12k 12k 12k 1 12k 3k, 12k 12k 6k 12k 6k 6k 3k, 6k 6k 6k, 12k 48k, 320k Filament diameter (mm) 7.0 5.1 5.1 6.5 Surface area (m2 g71) 0.45 0.5 0.5 0.45 Tensile strength (MPa) 3650 5650 4620 2900 3930 4138 5379 3447 3530 4900 5490 4220 3000 3200 3950 4800 5500 4500 3920 4810 5790 3240 3430 3330 3530 4210 4900 5490 6370 4410 4120 3820 3430 2740 3600 1552 Tensile modulus (GPa) 231 290 290 390 221 276 276 441 235 235 294 490 230 210 238 240 290 435 235 240 285 345 390 650 230 230 230 294 294 377 475 588 637 392 228 221 Tensile strain (%) 1.4 1.8 1.6 0.7 1.7 1.5 1.8 0.8 1.5 2.1 1.8 0.9 1.4 1.4 1.5 2.0 1.9 1.1 1.7 2.0 2.0 0.9 0.9 0.5 1.5 1.8 2.1 1.9 2.2 1.2 0.8 0.7 0.5 0.7 Density (g cm73) 1.76 1.81 1.78 1.81 1.79 1.73 1.77 1.87 1.79 1.80 1.80 1.88 1.78 1.8 1.77 1.8 1.8 1.81 1.77 1.78 1.80 1.79 1.85 1.97 1.76 1.78 1.80 1.81 1.80 1.77 1.88 1.94 1.98 1.81 1.78 1.75

Amoco [Thornel] Hexcel

Mitsubishi rayon [Pyrofil] Sigrafil Tenax

Toho rayon [Besfight]

Toray [Torayca] * available from Soficar

Zoltek [Panex]

6.8 7.0 7.0 7.0 5.0 4.7 7.0 7.0 5.0 6.7 6.2 4.1 7.0 7.0 7.0 5.0 5.0 5.0 5.0 5.0 5.0 7.0 7.4

Norita et al. (1988) have found that a new series of Torayca fibers exhibit higher compressive strengths than a conventional series of Torayca fibers with the same tensile modulus, though the new CFs have thinner diameters than those of the conventional fibers. They have described that the new CFs have a denser microstructure consisting of smaller and less oriented graphitic crystallites than the conventional CFs with the same modulus. During the creep test for which a Hercules HM3000 fiber yarn is strained to 140% at 2310 8C, the interlayer spacing reduces from 3.427 to 3.406 A, the crystallite size, Lc, in creases from 45 to 57 A, and the degree of preferred orientation increases. The postcreep fiber, furthermore, exhibits faint (112) reflections, though the as-received fiber does not exhibit them (Kogure et al., 1994).

1.01.3.4

Radial Heterogeneity

Within a carbon fiber filament there is a property or structure gradient across the radius of the filament. Some CFs exhibit a profound skin-core difference, with more highly oriented and larger crystallites in the skin region, and less oriented and smaller crystallites in the core. Diefendorf and Tokarsky (1975), Chen and Diefendorf (1872) (on Hercules HMS and HMU), Morita et al. (1977) (on Torayca T300 and two fibers with higher strength), and Sawada and Shindo (1981) (on Torayca T300 and Hercules HMS) found that there was a modulus gradient exhibiting a higher modulus in the outer skin region. Sawada and Shindo (1981) also found the radial gradient of density being lower in the inner region. Guigon and Oberlin (1986) showed that the crystallites in the surface

Structure
CARBON FIBER SKIN/CORE STRUCTURE
25 T50

13

20

HMS4

15 P75 10

0.0

0.2

0.4

0.6

0.8

1.0

Distance from Fibre Core, x/R

Figure 11 Variation of a crystallite orientation parameter, yh, across the fiber determined from the SAD patterns (reproduced by permission of Elsevier Science Ltd. from Carbon, 1995, 33, 97 107).

region of PAN-based fibers were much larger than those in the center of the fibers. The above-mentioned results were obtained by testing the CFs etched chemically, except by Diefendorf and Tokarsky (1975). Raman spectroscopy can also be a useful technique for studying the variation of microstructure across a CF filament (Morita et al., 1986; Huang and Young, 1995). Two firstorder Raman bands have been found in various types of graphite and CFs (Tuinstra and Koenig, 1970a, 1970b; Morita et al., 1986), namely 1580 cm71 and 1360 cm71 bands. The Raman band at 1580 cm71 has been attributed to the CC in-plane stretching mode of the graphite planes of an infinite crystal. The 1360 cm71 band is thought to be due to the crystal boundaries of graphite. The ratio of intensities of two bands, I1360/I1580, can be related to the crystal size, La (Tuinstra and Koenig, 1970a). Furthermore, the bandwidth reflects the orientation of the graphite layer plane with respect to the fiber axis (Katagiri et al., 1988). Huang and Young (1995) thus found from the radial variation of the ratio of intensities I1360/I1580 obtained from the Raman spectra of longitudinally sectioned fibers that the crystallite size decreases going from the core to the skin in HMS4 fiber. The half-width at halfmaximum intensity of the 1580 cm71 band for the HMS4 fiber also decreased with increasing distance from the core of the fiber to the skin, as shown in Figure 11. This implies that there is a highly oriented skin in HMS4 fiber. Katagiri et al. (1988) measured the I1360/I1580 values in the cross-section of a 2500 8C PANbased fiber, and set forth that the degree of graphitization of the fiber were higher both in

the skin and center regions. The profiles of the full-width at half-maximum of the 1580 cm71 band for the fiber also supported the result obtained from the band intensity ratios. Honjo and Shindo (1986c) anodically oxidized Torayca fibers T300 and M40 in an aqueous solution of sulfuric acid, and found that the filament skin was converted to a transparent thin-walled tube of graphite oxide, and the filament core was left unoxidized. Furthermore, other filaments of the T300 and M40 fibers were thinned by anodic oxidation in an alkaline solution, and then oxidized anodically in an aqueous solution of sulfuric acid. The thinned T300 filaments were tapered without forming graphite oxide, while the thinned M40 filaments converted to graphite oxide. Such phenomena seem to correspond to the structural variation along the fiber radius indicated by Katagiri et al. (1988). Huang and Young (1995) took TEM micrographs and corresponding SAD patterns from skin and core regions for CFs T50 and HMS4, and estimated the orientation parameters, i.e., half-width at half-maximum intensity of the (002) arc of the SAD patterns. From those values, they found that the orientation parameter decreases from 21.78 in the core to 118 in the skin for T50 and from 17.28 in the core to 14.78 in the skin for HMS4. In longitudinal sections of some high-modulus CFs, a continuum decrease of the transverse radius of curvature of lamellae from the external surface of the fiber toward the center was observed by Oberlin and Guigon (1988). In transverse sections they also observed the presence of a skin-core texture difference in (002) lattice-fringe images for some HM CFs (having the largest transverse radius of curvature of lamellae). That is, the majority of graphitic layers are parallel to the surface, whereas in the center the majority of layers are strongly misoriented.

Orientation Parameter (ho)

1.01.3.5

Schematic Structure

Several models of the structures of PANbased low- or high-modulus CFs have been proposed during the course of structural studies in the past (Shindo, 1961; Crawford and Johnson, 1971; Fourdeux et al., 1971; Diefendorf and Tokarsky, 1975; Bennet and Johnson, 1978; Guigon et al., 1984). They suggest a longitudinal preferred orientaion, sheathcore arrangement in longitudinal and transverse sections, undulation, and entanglement of stacks or lamellae of graphitic layers, and the presence of voids. Figure 12 shows a 002 lattice

14

Polyacrylonitrile (PAN)-based Carbon Fibers

Figure 12 002 Lattice fringe micrograph of CF M50J, cross-section (reproduced by permission of Elsevier Science Ltd. from Carbon, 1992, 30, 981987).

fringe micrograph of CF M50J, cross-section (Deurbergue and Oberlin, 1992). Figures 13 and 14 show three-dimensional structural models proposed by Bennet and Johnson (1978) and Guigon et al. (1984), respectively. In the Bennet and Johnson model, lamellae are entangled with each other longitudinally, a part of a lamella belonging to other lamellae and folded parallel to the fiber axis being caught with their neighbors transversely. In the skin region there are stacks of layers oriented circumferentially. Such a texture with preferred orientation has, in fact, been observed on a lattice-fringe image (Bennet and Johnson, 1979). Guigon and co-worker's model for highstrength CFs is made of a set of isometric lamellae, connected edge-to-edge, which fold parallel to the fiber axis. The structural unit in this model is a stack of aromatic layers with an extension of about 10 A on each edge. These units connect to each other to form columns along the fiber axis, though the units are in various orientations of twist both within and outside the plane of the unit. In the model for HM CFs, the structural units are larger in size and higher in degree of orientation.

1.01.4.1 1.01.4.1.1

Longitudinal Elasticity

PAN-based CFs are elastic, but their stress strain behaviour is nonlinear (Curtis et al., 1968; Jones and Johnson 1971; Hughes, 1986). Nukushina et al. (1986, 1989) have observed for several types of PAN-based CFs that the stressstrain curve of a CF filament obtained after the loading of 80% of a fracture stress and unloading had been repeated several times is the same as that obtained before the loading unloading cycle. Hayakawa et al. (1990) found that tangent and sonic moduli for CFs monotonously increased with increasing applied stress. Okada et al. (1995) have found that the preferred orientation of crystallites of several kinds of CFs with different HTTs increases linearly as

1.01.3.6

Chemical Composition

Nitrogen contents of CFs Torayca T300 and T900 are 6.3 and 5.4%, respectively. Oxygen contents of those fibers are 0.9 and 0.8%, respectively. Their hydrogen contents are below 1% (Trinquecoste et al., 1996).
Figure 13 Schematic three-dimensional representation of structure in high-modulus PAN-based CF. The layer planes are highly interlinked in both longitudinal and transverse direction (reproduced by permission of Kluwer Academic Publishers from J. Mater. Sci., 1983, 18, 33373347).

1.01.4

MECHANICAL PROPERTIES

Mechanical properties for various types of commercial CFs extracted from manufacturer's data sheets are shown in Table 2.

Mechanical Properties

15

Figure 14

(a) Model of high-modulus PAN-based CF. (b) Model of high tensile strength PAN-based CF (reproduced by permission of Elsevier Science Ltd. from Carbon, 1992, 30, 981987).

the strain rises up to about 1.3% on loading, but goes back to the original on unloading. The variation in scattering peak width of SAXS, giving information on the width of voids, parallel to the fiber axis with tensile stress also was almost reversible. Curtis et al. (1968) have found that the dynamic tensile modulus of HT and HM CFs increases markedly with increasing tensile stress, and that the increase in dynamic modulus accompanies an appreciable increase in crystallite orientation. Shioya and co-workers (Shioya and Takaku, 1994; Shioya et al., 1996) have measured the X-ray diffraction of CFs under tensile stress and have found that the preferred orientation increases with increasing stress. Furthermore, they found that the fiber compliance and the rotational compliance as its component decrease with increasing initial orientation, though the extensional compliance as the other component at smaller tensile stress is almost constant in the orientation range covering most of the CFs currently available.

1.01.4.1.2

Tensile modulus

The Young's modulus of CFs is considered to be determined by the volume (size) and modulis and preferred orientation, relative to the fiber axis, of aromatic layers (graphitic layer planes) or crystallites, and the volume and modulus of disordered regions, including cross-linking. The volume of the disordered regions is considered to be very small compared with that of the crystallites, and also those

moduli would be lower than those of the crystallites. Thus, Young's modulus of the CFs can be described as a function of size of the crystallites (in particular, La) (Oberlin and Guigon, 1988). Indeed, a good correlation has been found between Young's modulus and layer diameter for HM-type CFs by Guigon and Oberlin 1984). In CFs, the degree of preferred orientation of graphitic layer planes along the fiber axis also controls Young's modulus, and in addition the degree of orientation as well as crystallite sizes increase together as the HTT in the preparation rises. Young's modulus, therefore, can also be almost well expressed as a function of the degree of preferred orientation (Ruland, 1969; Fourdeux et al., 1971; Northolt et al., 1991). Ruland (1969) interpreted the Young's modulus as a function of orientation of graphitic crystallites composed of wavy graphitic layer planes by means of an elastic unwrinkling model using the intrinsic elastic constants of crystallites. Brydges et al. (1969) showed that the tensile moduli of CFs were rather close to values calculated on the basis of a polycrystalline model comprising perfect graphite crystallites with a measured amount of crystallite orientation by assuming uniform stress on crystallites. On the basis of the uniform stress model, Northolt et al. (1991) have also proposed an equation interpreting the modulus of CFs as a function of the shear modulus and orientation of graphitic crystallites. Okada et al. (1995) have pointed out that the observed values of tensile modulus as a function of tensile strain are markedly different from the values calculated based on the uniform stress model.

16

Polyacrylonitrile (PAN)-based Carbon Fibers


Table 3a Statistical values of tensile strength for CFs. Fiber T-300 M-40 Weibull Scale Location Mean strength modulus, m parameter s0 parameter sp at 30 mm length Length of links (GPa) (GPa) (GPa) 5.28 5.35 5.02 3.99 0.502 0.800 3.01 2.36 0.714 0.286

Source: Miwa et al. (1991).

Table 3b Statistical values of tensile strength for CFs. Fiber T-300 Sample length (mm) 30 Weibull modulus, m 5.5 Weibull scale parameter sL (GPa) 3.2 Mean strength at 30 mm length (GPa) 2.95

Source: R'Mili et al. (1996).

In general, the Raman bands shift to lower frequencies under tensile loading and to higher frequencies in compression. In most cases, an approximately linear relationship exists between the Raman band shift and the strain on fiber. The Raman band shift per strain increases proportionally with fiber tensile modulus (Huang and Young, 1995).

1.01.4.1.3

Tensile strength

At the present time, the ratio of tensile strength to Young's modulus for the commercial CFs in Table 2 is in the range 0.0050.022. Although the ratio for a graphite whisker (Bacon, 1960) is considered to be 0.02, the values for many brittle materials are in the range 0.10.2 (Johnson, 1987). Torayca T1000G (Yamane et al., 1987) having the highest strength, 6.4 GPa, among commercial CFs, exhibits a strength-to-modulus ratio of 0.022. Such low values of strength-to-modulus ratio are attributed to the defective nature of CFs. With the CFs produced earlier, flaws have been observed on fracture surfaces (Johnson, 1969; Sharp and Burnay, 1971; Moreton and Watt, 1974). Reynolds and Moreton (1980) have reported that strength-limiting flaws develop in CFs from impurity particles, particularly after heat treatment to 2500 8C. Even in the CFs produced later, the inhomogeneity in microstructure has been observed under TEM, which would act as a strength-limiting element (Dobb et al., 1995). The strength of CFs, thus, is gauge-length dependent owing to a random distribution of flaws or defects (Moreton, 1969). The strength

at a short gauge length as the intrinsic strength of CFs, then, was evaluated to obtain a realistic target to aim at in the pursuit of practical strength improvement (Diefendorf and Tokarsky, 1975). The tensile strength values of CFs at very short gauge lengths, i.e., critical lengths, were also necessary for determination of the fibermatrix interfacial properties in CFRPs by the fragmentation method (Asloun et al., 1989; Hitchon et al., 1979). Such strengths were estimated by means of Weibull analysis and extrapolation by assuming a linear logarithmic dependence on gauge length of tensile strength. Table 3 shows the tensile strength and Weibull parameters for Torayca fibers estimated by Miwa et al. (1991) and R'Mili et al. (1996). The statistical variation in mechanical properties of T300 fiber has been studied by using monofilaments and loose bundles of 6000 filaments, and the strength distribution is analyzed by the twoor three-parameter Weibull model (R'Mili et al. 1996). The mean tensile strength of CFs in a bundle can be determined by treating the filament-strength distribution from the stressstrain curve of the bundle with Weibull statistics (Noguchi et al., 1976; R'Mili et al., 1996). Chi et al. (1984) also determined the filament strength distribution from fiber bundle testing. Fiber-resin transfer lengths of 0.5 mm (Hitchen and Phillips, 1979) and 0.6 mm (Noguchi et al., 1976) have been obtained, and about 5, 8.5, and 10 GPa strength at that gauge length have been given to Torayca T300, T800H, and T1000, respectively. A method for determining the bending strength at a very short gauge length is the elastica test (Jones and Duncan, 1971). Table 4 gives the results of the loop test for some CFs

Mechanical Properties
Table 4 Mechanical properties of CFs from loop tests at RT. T300 Loop test straina Tensile modulusb Tensile strengthb Loop test strengthc % GPa GPa GPa 2.78 230 3.50 6.4 T700S 5.04 230 4.80 11.6 T900 4.18 294 5.40 12.3 M40 1.38 392 2.70 5.4

17

a Measured. bFrom literature. cCalculated. Reproduced by permission of Elsevier Science Ltd. from Trinquecoste et al. (1996).

(Trinquecoste et al., 1996). In the loop test, the shape of the stressed loop showed a very good superposition with the theoretical curve (called elastica). The bending strength is obtained according to the formula st = Ee based on a perfect elastic material and assuming that the tensile and compressive moduli are identical. Compared to a regular single-fiber tensile test, the working volume of the loop-tested sample is much smaller. The loop test strength values in the table, hence, are twice the commercial values. It is well known that the strength of many kinds of fibers increases with decreasing fiber diameter, and such a behavior has been also observed for CFs (Shindo, 1961; Jones, 1971; Jones and Duncan, 1971). Figure 15 shows the tensile strengths at a gauge length of 0.1 mm for Besfight and Torayca high-strength CFs as a function of cross-sectional area (Shindo and Sawada, unpublished). These fibers were selected as CFs having Young's moduli, 240 250 GPa, being close to each other. Those strength values were obtained by Weibull analysis of strengths of about 40 filaments of gauge lengths 5, 10, 20, and 40 mm and extrapolation. It can be seen in Figure 15 that the smaller the fiber diameter, the higher the tensile strength becomes. As mentioned in Section 1.01.3.4, many types of CFs have radial heterogeneity. Accordingly, it is considered that the relative volume of the skin region increases and the structure of CFs becomes more homogeneous as the cross-sectional area decreases. This is considered to be one of the causes of the higher strength of the thinner CF. Extrapolation of the line of least squares in the figure to a fiber diameter of 1 mm gives a strength of 16.8 GPa. Although the slightly lower value seems much more likely for the strength in this case, such a strength may suggest the tensile strength of fracture at a disordered region or cross-linking between crystallites along the fiber axis in the skin region of the CFs. The tensile strength of PAN-based CFs increases with HTT, reaches a maximum at around 15001600 8C (depending to operating conditions), and finally decreases more slowly

at an HTT higher than 2500 8C (Figure 6). The decrease in tensile strength is attributed to a decrease in the amount of disordered region linking the ordered or crystalline regions (Johnson and Tyson, 1970; Mittal et al., 1997a). As mentioned in Section 1.01.3.3, high-strength type and intermediate-modulus type CFs exhibited rough, rather poorly defined granular texture in high-resolution SEM images of tensile fracture surfaces. Thus, it is considered that the initiation of a fracture in such fibers probably occurs in longitudinal bonds of a disordered region between two crystallites, its propagation being transversal. On the other hand, some HM CFs showed fracture surfaces exhibiting sheet-like textures. In such CFs, tensile fracture may be initiated at a flaw on lamella or a pinched-in region of lamella. The initiation at a flaw on lamella has been explained with the mechanism proposed by Ben-

Figure 15 Tensile strength at a gauge length of 0.1 mm for Besfight and Torayca high-strength CFs as a function of cross sectional area. B: Besfight. T: Torayca. Subscripts of u and s stand for the fibers without and with surface treatment, respectively. The line in the figure represents an equation of least squares, y = 16.6140.2766x, where y and x are tensile strength and cross-sectional area, respectively (after Shindo and Sawada, unpublished).

18

Polyacrylonitrile (PAN)-based Carbon Fibers ken fiber fragment length, loop, and fiber encapsulated into block tests have been used for determining the compressive strength of CFs (Kozey et al., 1995). The single filament recoil technique is applicable to fibers with a lower compressive strength than tensile strength (Allen, 1987). In the broken fiber fragment length method, axial tensile or compressive stress is applied to a filament by bending a rectangular resin beam in which the filament is embedded near the beam surface. The mean tensile strength of the fiber at a length, and mean fragment length in tension and compression, are used in the calculation of compressive strength (Ohsawa et al., 1990; Miwa et al., 1991). Compressive strengths for various types of CFs are shown in Table 5. These compressive strength values were determined from the compressive strength of composites by normalizing to 100% fiber (Kumar et al., 1993). As seen there, the compressive strengths even for the same fiber are considerably different according to the test method or author. The same tendency can also be seen in a table by Peebles (1994). The results of compression tests depend strongly on the loading geometry and test conditions. Methods such as the bending beam and elastica loop tests have, by design, stress fields that are not pure axial compression. Tests with embedded fibers (or a single fiber) in a matrix are liable to be affected by the residual stress on the tested fibers caused by the matrix shrinkage, and the misalignment of the fibers during the test (Jiang et al., 1993). Osawa et al. (1990) and Miwa et al. (1991) have estimated the compressive strength when the thermal stress working from resin perpendicularly to the fiberresin interface could be regarded as zero. The recoil strengths of PAN-based CFs, shown in Table 6 (Dobb et al., 1995), however,

Figure 16 Reynolds and Sharp mechanism of tensile failure. (a) Misoriented crystallite linking two crystallites parallel to the fiber axis. (b) Tensile stress exerted parallel to fiber axis causes layer plane rupture in direction La\, crack develops along La\ and Lc. (c) Further exertion of stress causes complete failure of crystallite. Catastrophic failure occurs if the crack exceeds the critical size in Lc or La\ directions (reproduced by permission of Kluwer Academic Publishers from J. Mater. Sci., 1983, 18, 33373347).

nett and Johnson (1983) using the Reynolds and Sharp (1974) criterion. Their proposal is based on the fact that the internal and surface flaws that initiated failure showed evidence of large misoriented crystallites in the walls of holes. According to Reynolds and Sharp, fracture of a CF is caused by basal plane rupture in crystallites at large misorientation angles. Their mechanism is explained in Figure 16.

1.01.4.1.4

Compressive strength

Several test methods such as tensile recoil (Allen, 1987), unidirectional composite, bro-

Table 5 Compressive properties of CFs. Commercial name T-300a AS-4c T-40a T-1000d IM8c T-50a M40Jd GY-7b M-60Jd Tensile modulus (GPa) 235 235 290 295 310 390 390 520 585 Tensile strength (st) (GPa) 3.2 3.6 5.7 7.1 5.17 2.4 4.4 1.8 3.8 Compressive strength (sc) (GPa) 2.88 2.69 2.76 2.76 3.22 1.61 2.33 1.06 1.67 scst71 Density (g cm73) 1.76 1.80 1.81 1.82 1.80 1.81 1.77 1.96 1.94

0.9 0.75 0.48 0.39 0.62 0.67 0.53 0.59 0.44

Source: Kumar et al. (1993). a Amoco Thornel; bBASF Celion; cHercules Magnamite; dToray Torayca.

Mechanical Properties
Table 6 Mechanical properties of various CFs (single fiber test). Commercial name Fiber type Tensile (st) Modulus Strength Strain (GPa) (GPa) (%) HM-S Grafila S-g Besfightb C1 C2 X Y Z T1000 320 220 370 330 280 255 2.1 3.0 3.4 3.1 3.5 5.7 0.36 1.35 0.93 0.96 1.25 1.87 Compressive (sc) Strength (GPa) 0.8 1.5 1.4 1.3 1.9 2.2 Strain (%) 0.26 0.69 0.38 0.40 0.67 0.87 0.38 0.50 0.41 0.42 0.54 0.39 scst71

19

Toraycac

Source: Dobb et al. (1995). a Coutaulds; bToho rayon; cToray.

Table 7 Commercial name T-1000a IM7b T650/35c G30500d Grafil 33500 AS4b

Compressive properties of CFs. Compressive strength, sc (GPa) 9.9 8.1 7.7 8.3 9.0 8.5 Compressive strain (%) 73.3 72.9 73.2 73.6 73.9 73.7 scst71 1.41 1.51 1.69 2.19 2.01 2.13

Tensile strength, st (GPa) 7.03 5.38 4.55 3.79 4.48 4.00

Source: DeTeresa (1991). a Toray Torayca; bHercules Magnamite; cAmoco Thornel; dBASF Celion.

are typically lower than those obtained from composite tests. Jiang et al. (1993) proposed a universal logistic model which provides a good fit to the recoil strength distributions obtained at different gauge lengths. This analytical expression can be used to obtain a physically meaningful extrapolated average zero gauge length recoil strength. As shown in Table 7, the failure strains of CFs obtained in the compression test utilizing the piezoresistivity by which the axial compression strain of the filaments was accompanied were 73% or greater (DeTeresa, 1991). Any compressive strength calculated assuming linear elastic behavior will most likely be an overestimate. These compressive strength values, nevertheless, are markedly higher than those obtained by the composite and recoil methods. The compressive strengths calculated from the strains obtained for CFs embedded in resin by Prandy and Hahn (1991) are also fairly high. As regards the fracture mechanism, it has been reported that PAN-based CFs, which are thin and strong compared with pitch-based CFs, exhibit buckling and/or kinking fracture (Dobb et al., 1990; Prandy and Hahn, 1991; Kumar et al., 1993); such failure modes are in

contrast to that of more crystalline CFs, such as pitch-based HM CFs, having sheet-like microstructures, which tend to fail in transverse shear mode. However, there is a report which concludes that the single filaments of PAN-based CFs, AS4, IM6, and IM7, embedded in an epoxy polymer, fail in transverse shear and not in a microbuckling mode (Boll et al., 1990). The compressive strength of PAN-based CFs increases with increasing tensile strength (Dobb et al., 1990) and (002) d-spacing, and with decreasing tensile modulus, orientation parameter, crystallite sizes, and void content, respectively (Masson and Bourgain, 1992; Kumar et al., 1993; Dobb et al., 1995; Norita et al., 1988; Sumida et al., 1989). With boron-ion implantation, the compressive strength and torsional modulus of CFs (Table 8) increase by up to 25% and 50%, respectively, while the crystallite size, Lc, decreases (Matsuhisa et al., 1991). A correlation between torsional modulus and compressive strength was reported for PAN-based CFs and others (Norita et al., 1988). Dobb and co-workers, furthermore, have observed that the compressive strength of PAN-based CFs linearly increases with increas-

20

Polyacrylonitrile (PAN)-based Carbon Fibers


Table 8 Effect of B+ dose on compressive strength, torsional modulus, and tensile strength of single filament. Property Fiber 0 Compressive strength (GPa) Torsional modulus (GPa) Tensile strength (GPa) T1000G M40 T1000G M40 T1000G M40 3.63 21.60 14.70 6.18 3.33 B+ dose (ions cm72) 1015 5.00 20.60 4.02 1016 7.16 31.40 27.40 7.45 4.21 1017 7.84 7.74 25.50 4.31

Source: Matsuhisa et al. (1998).

ing intercrystalline disorder or intracrystallite disorder, along different lines for the group consisting of newly developed fibers, which have much improved compressive strength, and the conventional fibers group (Dobb et al., 1995). Norita et al. (1988) and Sumida et al. (1989) have reported that Torayca H and J series fibers exhibit fairly high compressive strengths as well as tensile strength compared with the conventional high-modulus fibers. It would appear from Figure 17 that the ability of CFs to withstand compressive stress is directly proportional to the degree of intercrystallite disorder. The same tendency can also be seen for the degree of intracrystallite disorder. According to Dobb et al., although all PAN-based CFs consist of two major phases (i.e., crystallites and disordered material), there are additional differences, depending on the fiber types, in (i) the distribution of crystallites and disordered material within the cross-section, and (ii) the distribution in the sizes and the orientation of crystallites across the cross-section. The uniformity of the size distribution suggests that compressive fracture is not governed by a random flaw distribution as in the case of tensile fracture (Dobb et al., 1995). More likely, failure is determined by some microcrystalline structure that is uniformly distributed along the fiber (Boll et al., 1990). Although the total amount of disordered regions plays an important role in determining the compressive strength and failure mechanisms, it is equally important that the disordered region is homogeneously distributed throughout the fiber, as shown in Figure 18. According to Dobb et al. (1995), it is also important that crystallites have dimensions below about 5 nm in all directions, and that their preferred orientation is maintained. Norita et al. (1988) proposed the ratio of tensile modulus to shear modulus as a measure of anisotropy in various fibers, and found that the compressive strength of unidirectional CF

composites decreases with increasing ratio and with increasing CF tensile modulus. Strong transverse cohesive force, like covalent crosslinks, are considered to increase the macroscopic shear modulus and to improve the compressive properties (Northolt et al., 1991).

1.01.4.2 1.01.4.2.1

Transverse Transverse modulus

Hayakawa et al. (1994) found that PANbased CFs can be deformed reversibly by

Figure 17 Compressive failure stress of CFs vs. intercrystallite disorder Dd (reproduced by permission of Elsevier Science Ltd. from Carbon, 1995, 33, 15531559).

Mechanical Properties

21

Figure 18 Lattice-fringe TEM image of CF T1000. Marker bar: 6 nm (reproduced by permission of Elsevier Science Ltd. from Carbon, 1995, 33, 15531559).

transverse compression up to a displacement corresponding to about 10% of the diameter, the compressive deformation being more difficult with increasing compressive deformation, and that the transverse modulus decreases with increasing crystallite orientation or longitudinal modulus. As shown in Table 9, the fibers with longitudinal moduli from 226 to 581 GPa exhibited transverse moduli from 13.9 to 6.9 GPa.

1.01.4.2.2

Torsional modulus

The torsional modulus of PAN-based CFs decreases with increasing Young's modulus. The torsional strength of the fibers increases as the torsional modulus increases. The torsional strength also increases with increasing tensile strength (Table 10) (Sawada and Shindo, 1992; Norita et al., 1988; Sumida et al., 1989). It is considered that the torsional modulus depends on the size of the graphitic layers or crystallites, their amount, their degree in radial and axial orientation, and the amount of disordered region, including cross-linking, between the layers or crystallites. It is also considered here that the transverse cohesive force, depending on the amount of disordered region, plays an important role in increasing the torsional modulus and strength. Figure 19 shows a torsional fracture surface of CF Besfight STA. The shape suggests that the transverse cohesive force is high compared with that

Figure 19 A torsional fracture surface of CF Besfight STA, torsional modulus: 24.0 GPa, Young's modulus: 263 GPa, torsional strength: 1.28 GPa, tensile strength: 4.04 GPa. Views from (a) top and (b) a side (reproduced by permission of Elsevier Science Ltd. from Carbon, 1992, 30, 619 629).

22

Polyacrylonitrile (PAN)-based Carbon Fibers


Table 9 Fiber Transverse compressive properties of CFs. Orientation Longitudinal Transverse Compres. parameter modulus modulus strength (17 bp71) (GPa) (GPa) (GPa) 0.808 0.830 0.887 0.915 0.945 0.779 0.901 0.946 226 267 379 469 581 250 398 588 13.9 12.7 11.6 9.1 6.9 16.8 6.8 5.5 (1.9) (0.4) (1.1) (2.6) (1.1) 4.23 1.79 1.57

Diameter Density (mm) (g cm73) 7.4 5.8 5.4 5.3 5.0 6.73 6.7 4.2 1.77 1.84 1.83 1.90 2.03

T03 T10 M40 M50 M60 Besfight STIII HM40 HMS6X

Source: Hayakawa et al. (1994); Besfight: Fujita et al. (1992). b, Width at half-maximum intensity of (002) peak. Standard deviation.

Table 10 Torsional properties of CFs. Fiber Torayca an early CFa Exp. type IIIa Besfight HTa STAa GIRIO Exp. CFa S H30a Magnamite ICMa 24/3Aa Torayca M40a Besfight HMa Torayca T300b M30b M40b M46b T800b M40Jb Young's modulus (GPa) 201.4 269.1 235.9 262.9 273.9 235.1 441.7 431.4 558.4 466.3 230 294 392 451 294 392 Torsional modulus (GPa) 23.72 23.61 25.03 24.03 23.82 25.08 21.85 21.47 22.55 23.03 16.7 16.7 15.7 14.7 17.6 18.7 Tensile strength (GPa) 2.316 4.154 2.839 4.044 3.028 2.474 2.327 2.167 2.607 2.833 Torsional strength (GPa) 1.000 1.052 1.379 1.279 0.965 1.106 0.756 0.740 0.796 1.113

Source: aSawada and Shindo (1992); bNorita et al. (1988); Toray (private communication, 1998). GIRIO: Government Industrial Research Institute at Osaka. S: a company in Japan.

of high-modulus pitch-based CFs. The ditches running along the fiber axis should be effective for diminishing the torsional strength, as the stress tends to concentrate at the bottom of the ditches (Sawada and Shindo, 1992). Brydges et al. (1969) obtained a torsional shear modulus of 24.1 GPa for a HM fiber. Villeneuve et al. (1993) have shown the variations of the radial strain as a function of the axial strain, both measured directly on the TEM images recorded under axial tensile loading. A Torayca T300 filament with 6.2 mm diameter exhibited a radial contraction as the axial load was progressively increased, though a T300 filament with 6.9 mm diameter exhibited expansion. Villeneuve et al. (1993a) have tentatively calculated a Poisson ratio at room temperature from TEM measurements for the

T300 fiber with a 6.2 mm diameter. The value is 0.22, whereas the values in the literature are in the range 0.240.35. Krucinska and Stypka (1991) have reported a Poissons' ratio of 0.28 for Safil.

1.01.4.3

High-temperature Properties

CFs exhibit creep behavior under tension at high temperatures. Sines et al. (1989) carried out creep tests on HM3000 yarn specimens (Hercules, 3000 filaments, with a nominal diameter of 7 mm), applying stresses ranging from 455 to 648 GPa at temperatures from 2120 to 2430 8C for several hours. Elongation up to 140% was measured on these fibers. For

Electric and Magnetic Properties


Table 11 Physical properties of selected types of CFs in manufacturers' data sheet. Mfr. Amoco [Thornel] Fiber type

23

Carbon Tensile Specific Thermal Electrical CTE content modulus resistivity heat conductivity (%) (GPa) (1073O cm) (cal. g71 8C71) (1076 8C71) (cal. cm71 s71 8C71) 92 94 94 99 231 290 290 390 238 290 235 295 345 380 230 230 250 230 294 294 1.8 1.45 1.42 0.95 1.6 1.45 1.5 1.4 1.0 0.9 1.7 1.5 1.6 1.6 1.4 1.4 70.6 70.75 70.75 71.13 70.1

T-300 T-40 T650/42 T-50 Tenax HTA IMS Toho rayon HTA [Besfight] IM400 HM35 HM40 Toray T300 [Torayca] T300J T400H T700S T800H T1000G

0.17

4.1 6 1072

93 94 94 93 96 95

0.19 0.18 0.18 0.18 0.18 0.18

70.41 70.43 70.45 70.38 70.56 70.55

2.5 6 1072 2.23 6 1072 2.52 6 1072 2.24 6 1072 8.39 6 1072 7.65 6 1072

steady-state creep the apparent activation energy and activation volume were 1058 kJ mol71 and 2088 A3, respectively. These values of apparent activation energy and apparent activation volume suggest that the creep occurs on a crystallite scale. An apparent non-Hookean behavior has been observed on some Torayca CFs at temperatures above 1000 8C during thermal expansion experiments, the heated fibers keeping the memory of their last shape. A similar observation was also made in high-temperature loop tests. Over 850 8C, a plastic deformation begins to occur (Trinquecoste et al., 1996). In this temperature range, a fairly large amount of nitrogen is eliminated from the fiber. Accordingly, such an apparent inelastic behavior seems to be induced by a structural rearrangement in the fiber when heteroatoms, mainly nitrogen, are driven out. For example, CFs T300 and T900 exhibit nitrogen contents of 6.3% and 5.4%, respectively, but after heat treatment to 2000 8C, both the fibers contain no nitrogen.

power at temperatures from 0 to 300 K for Torayca intermediate (TH series) (T800H, T-1000G) and HM (MJ series) type (M40J, M46J, M50J, M60J) fibers. According to their results, the M-type (J-series) fibers indicate the semiconductor-like temperature dependence all over the temperature ranges examined, and the resistivity increases with decreasing elastic modulus. On the contrary, the T-type fibers exhibit a peak in resistivity around 35 K, as shown in Figure 20. The metallic-like

1.01.5 1.01.5.1

ELECTRIC AND MAGNETIC PROPERTIES Electrical Resistance and Thermoelectric Power


Figure 20 Resistivity vs. temperature curves for Torayca T800H. Solid line represents the calculation (reproduced by permission of the authors from Matsubara et al., (1995).

The electrical resistance of CFs, Torayca, Besfight, and Amoco decrease with increasing Young's modulus. These values are shown in Table 11. Matsubara et al (1995) have measured electrical resistivity and thermoelectric

24

Polyacrylonitrile (PAN)-based Carbon Fibers linearity, and reaches a value of 0.0019 or 0.0021 at a strain of 1.1% under static tension. The strain sensitivity values obtained are 1.8 and 1.9; (ii) The strain and DR/R0 obtained during cyclic tension are totally reversible at low values of the stress amplitude (up to 58.1% of the fracture stress), but their irreversible components increase with stress amplitude at high values of the stress amplitude; (iii) The observed irreversible resistance change is attributed to damage, as supported by the accompanying decrease in the elastic modulus. Damage can occur even in the regime of elastic deformation; (iv) Irreversible strain occurs at stress amplitudes 573.1% of the fracture stress and increases with increasing stress amplitude; (v) The reversible resistance change is mainly due to the dimensional change associated with elastic deformation; (vi) The strain sensitivity (reversible DR/R0 per unit reversible strain) is 1.92.3 and is quite independent of stress or cycle number (from 1 to 2). These values are close to those obtained by Owston (1970). The CF embedded in epoxy resin exhibits the opposite behavior from that mentioned above. The embedded fiber showed the decrease of the fractional change in resistance upon tensile loading and its recovery upon subsequent unloading. The decrease in the fractional resistance change is attributed to the reduction of the residual compression stress in the fiber having been induced on curing and cooling of the resin (Wang and Chung, 1997b).

Figure 21 Temperature dependence of thermoelectric power for high modulus CFs (reproduced by permission of the authors from Matsubara et al., (1995).

temperature dependence observed in spite of their lower crystallite perfection compared with M-type fibers can be explained by considering the Rayleigh wave phonon whose velocity is so small that a number of phonons are excited even at liquid helium temperature. They explained the behavior assuming a mixture model of the band conduction and two-dimensional variable-range hopping conduction. Figure 21 shows the temperature dependence of the thermoelectric power of Torayca MJ-type fibers (Matsubara et al., 1995). The M-type fibers exhibited the positive temperature dependence of thermoelectric power (S), and the absolute values decrease with decreasing modulus. The T-type fibers showed the temperature dependence of thermoelectric power similar to that of the resistivity. Such a dependence implies that the wave phonons affect the thermoelectric power. 1.01.5.2 Electromechanical Properties

1.01.5.3

Magnetoresistance

Values of average crystallite layer plane transverse magnetoresistance for CFs are shown in Table 12 (Hishiyama et al., 1984). The CFs from PAN fiber without stretching during stabilization treatment exhibit negative values. The CFs prepared from PAN fiber stretched during stabilization treatment, however, exhibit positive values. Furthermore, it can be seen that the stretching at 2700 8C is effective for improving the degree of graphitization (Hishiyama et al., 1991).

Single filament electromechanical behavior for a Torayca T300 fiber has been observed by measuring the fractional increase in electrical resistance (DR/R0), and stress and strain simultaneously obtained during static and cyclic tension to failure. Wang and Chung (1997a) have found the following phenomena: (i) DR/R0 increases monotonically with strain stress, with a slight negative deviation from

1.01.6 1.01.6.1

THERMAL PROPERTIES Thermal Expansion

The coefficient of thermal expansion along the basal plane for single crystal graphite, aa, is 71.6 6 1076 K71 around RT, and then increases up to + 1.2 6 1076 K71 over 1000 8C. The coefficient of thermal expansion in the

Thermal Properties
Table 12 Average crystallite layer plane transverse magnetoresistance (Drr71)cr at liquid nitrogen temperature for CFs. HTT 3000 2900 2800 3000 2830 3000 2830 2700 3000 2830 2700 3000 2830 2700 Treatment (Drr71)cr 70.547 70.954 70.618 26.18 6.37 26.20 9.19 2.10 21.18 9.23 4.74 22.99 4.73 1.54

25

Stretcheda SG-4b SG-8b SG-12b

Figure 22 Coefficients of longitudinal thermal expansion of CF filaments (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996, 34, 923929).

Source: Hishiyama et al. (1984). a Stretched with a load of 40 mg per filament both in stabilization and carbonization.bPrepared from the CF with treatment of stretcheda, by stretching at 2700 8C with loads of 40 mg per filament (SG-4), 80 mg per filament (SG-8), and 120 mg per filament (SG-12).

direction perpendicular to the basal plane, ac, is & 27 6 1076 + 3 6 1079 K71 in the temperature range from RT to 800 8C (Trinquecoste et al., 1996). The coefficient of longitudinal thermal expansion for CFs decreases with increasing Young's modulus. This can be apparently shown when the CTE values for Torayca and Thornel fibers in Table 11 are plotted as a function of Young's modulus. The correlation between Young's modulus and CTE has been already observed by Wolff (1987); an abrupt change is present in the slope of the curve at about 680 GPa, where the plotted values are for single-crystal graphite and pitch-based CFs with higher Young's moduli than those of PAN-based CFs. With longitudinal thermal expansion, ak, of CFs, T300, T700S, and T900, as shown in Figure 22 (Trinquecoste et al., 1996), all of the fibers behave similarly on increasing temperature: slight contraction from RT to around 400 8C; expansion up to 2000 8C; burning at higher temperatures. The contraction in the first temperature region is comparable to the negative value of the aa coefficient of the graphite single crystal at low temperatures. Comparison of the ak coefficients shows that T300 and T700S fibers present a very similar behavior, while T900 has a much lower ak coefficient. The ak coefficients up to 1000 8C have also been measured on fibers T300 and M40 by Yasuda et al. (1987), obtaining a similar beha-

vior as by Trinquecoste et al. (1996). Villeneuve and Naislain (1993) measured longitudinal thermal expansion for T300 by TEM, and obtained its mean coefficients calculated for various temperature ranges, RT to 800 8C. According to Trinquecoste et al. (1996), the radial thermal expansion coefficient is 50 6 1076 K71 for Torayca T900, and that for M60J measured with a composite in the range 4090 8C by Nanjyo et al. (1994) is 15 6 1076 8C71. Villeneuve et al. (1993) obtained mean coefficients of radial expansion for T300 fiber, for various temperature ranges from 25 8C to 800 8C based on measurements by TEM, and found that it decreases from 17.6 6 1076 8C71 for a range of 25200 8C to 1.2 6 1076 8C71 for a range of 600800 8C.

1.01.6.2

Thermal Conductivity

The thermal conductivity of CFs increases with increasing Young's modulus. This can be seen if the values for Torayca fibers in Table 11 are plotted as a function of Young's modulus. Low thermal conductivity PAN-based CFs are used as reinforcements of phenolic resin matrix composites used as ablative insulators in solid rocket motor nozzles and exit cones. The CFs with HTTs of 900 8C, 1120 8C, and 1350 8C show thermal conductivities as low as 2.0, 4.0, and 13.6 W m71 K71, respectively (Katzman et al., 1994). After Heremans et al. (1985), the thermal conductivity of PAN-based CFs increases from the order of 1071 to the order of 102 W m71 K71 with increasing temperature from several tens to several hundreds of degrees Kelvin.

26 1.01.7 1.01.7.1

Polyacrylonitrile (PAN)-based Carbon Fibers SURFACE PROPERTIES Morphology and Surface Areas spectrum as a measure of structural deviation from the completeness of the graphite structure. The parameter ac is defined by the ratio a/b, where a is the long tailing component toward a higher binding energy, and b is the short component of FWHM (full-width at half-maximum), namely a+b, of the C1s spectrum. They reported that it increases in Torayca M40 by a surface oxidation treatment. This disordering was not restored after the fiber was subjected to a thermal treatment at 1000 8C in a vacuum, where surface functional groups are removed. 1.01.7.2 Functional Groups

The open pore volume (porosity) of CFs increases with increasing HTT up to around 1000 8C, and then decreases to be negligible at HTTs above 1200 8C (Spencer et al., 1970; Ismail, 1991). The surface of CFs has many wrinkles or ridges running nearly along the fiber axis. Surface roughness results from such wrinkles or ridges as well as depressions on the fiber surface. The ratios of BET surface areas to the geometrical areas estimated from laser diameters can be regarded as a measure of surface roughness. Table 13 shows those areas and surface roughness for some Torayca and Besfight CFs (Shindo, 1983). The Besfight fibers have a circular cross-section, but the Torayca fibers used have a cross-section of broad bean shape. The circumference of the cross-section for the Torayca fibers was then estimated with an approximation that the cross-section has a shape of an ellipse with apparent major and minor axes. Table 13 also shows another measure of geometrical surface areas obtained from the densities by a density gradient column test. Both of these values almost coincide with each other. Surface area of CFs can be determined by measuring the amount of a monomolecular layer of nitrogen or krypton on the CF surface using the BET method. ASA, which is the area occupied by chemically active sites on the fiber surface, can be determined by measuring the amount of oxygen chemisorbed at 300 8C on the surface of fiber degassed at 1000 8C in a high vacuum, under the assumption that the chemisorbed oxygen forms CO groups at active sites located at the basal plane edges of the crystallites. Table 13 also shows the active surface areas for the Torayca and Besfight CFs with and without surface treatment. Ismail (1987) obtained an ASA of 0.075 for an Amoco T300 fiber by extrapolating a plot of ASA vs. CO2 produced at 300 8C to zero CO2. A BET area for the fiber obtained by Ismail was 0.56 m2 g71. The ASA value is 13% of the BET SA value. The ASAs for the oxidized fibers in the table are in the range 913%. Those for the fibers without surface treatment are in the range 58% for HS type fibers, and 3% for HM type fibers. Such low values of ASA correspond to the fiber cuticle composed of graphitic basal planes parallel to the fiber surface. The surface treatment to which M-40 has been subjected seems to be severe compared with the others. Takahagi and Ishitani (1988) have proposed the asymmetric parameter, ac, of the XPS C1s

XPS is a very powerful technique for characterizing functional groups on CF surfaces. Oxidized Torayca CF T300 has markedly larger amounts of carbonyl and hydroxyl groups and a moderately larger amount of carboxyl groups than the unoxidized T300 fiber. The concentration of these groups was determined by XPS after the functional groups had been modified by chemical reactions with fluorine compounds. The relative amounts of oxygen estimated from the amounts of three functional groups determined by the surface modification reaction, however, were about 20% of the observed values from a direct measurement of O1s spectra for both the unoxidized and oxidized CFs. In their previous paper Takahagi and Ishitani (1984) concluded that hydroxyl, carbonyl, and carboxyl groups are major functional groups formed by the surface oxidation treatment, whereas the amounts of ester and ether groups are negligible. The chemical modification reagents may react only with the functional groups in the outermost layer (510 A) in the CF surface because of the poor permeability of reagents into CF. If there are more functional groups in a greater depth within the detection depth of XPS (3050 A), the direct O1s measurement will give a higher relative intensity (Takahagi and Ishitani 1988). Nakayama et al. (1990) determined the concentration of functional groups on the surface of weakly and strongly oxidized Torayca T800 fibers by curve-fitting the C1s spectrum using an asymmetric peak shape or a symmetric peak shape. Amounts of groups such as hydroxyl, carboxyl, and amine were estimated by gas chemical modification using fluorine compounds (Table 14). Consequently, the real surface of the oxidized CF was considered to be composed of both graphite-like and aliphatic structures with functional groups. Estimation of functional groups on CF surfaces using chemical modification were also

Table 13

Surface properties of CFs. Surface area of one filament (1076 m2 m71) evaluated from Density and Roughness Laser diameter (SAL) weight (SAD) SA (SAB) (SABSAL) 6 (SAL)71 (%)

Fiber

Increment Rate of ASA to Increment ASA Young's SA SA (%) of ASA by modulus (m2 g71) of SA by (m2 g71) STr (%) STr (%) (GPa)*

HS X550-Ua

HM M40-Ua

X550-Sa ST III-Ub ST III-Sb M40-Sa HM40-Ub HM40-Sb

244 240 250 247 392

0.603 0.630 0.473 0.488 0.526 1.199 0.496 0.592

4.47 3.17 128 19.3

0.046 0.085 0.025 0.049 0.014 0.138 0.013 0.052

7.7 13.4 5.2 10.0 2.7 11.5 2.6 8.8

74 92 325 238

19.5 19.4 21.1 21.1 20.4 20.7 20.8 20.8

19.5 19.4 21.1 21.3 19.6 20.2 20.2 20.6

32.1 33.4 29.8 31.4 31.1 70.2 31.0 37.1

65 73 41 49 53 238 49 78

Source: Shindo (1983). *Manufacturer's data, a Torayca,

Besfight, SA: BET surface area, ASA: Active surface area, STr: Surface treatment.

28
Table 14 Sample :COH

Polyacrylonitrile (PAN)-based Carbon Fibers


Surface chemical composition of surface-oxidized Torayca T800H fiber. O/C 6 103 7COOH :C-O-C: 7COOC: =C=O, etc. 30 75 244 :Si-OTotal 7NH2 N/C 6 103 =NH =N7 24 28 29 Total

Control Weakly oxidized Strongly oxidized

1 3

2 5 31

14

46 81 278

1 3 5

25 31 34

Reproduced by permission of Elsevier Science Ltd. from Nakayama et al., 1990.

Table 15 Fiber

Concentration of functional groups per hundred surface atoms, based on functional group derivatization and fluorine or mercury XPS analysis. C 81 85 82 95 85 84 O 14 11 15 2.7 11 14 N 5.2 2.0 2.4 1.9 4.1 1.6 OH 0.7 0.5 0.7 0.2 0.1 0.2 7C=O 1.1 1.4 1.3 1.8 0 0.2 COOH 1.8 1.6 1.2 0.6 0.4 0.3 OH+C=O +COOH 3.6a 3.5a 3.2a 2.6a 0.5b 0.7b

IM7c T65042d G40800e T40d AS-4c Celion 6000e

Source: aChan et al. (1991); bDeVilbiss and Wightman (1986). c Hercules, d Amoco. e BASE. C, O, and N: total carbon, oxygen, and nitrogen determined by XPS prior to derivatization.

applied to Hercules AS-4 and Celenese Celion 6000. Fluorine-containing compounds and a mercuric fluorine-containing compound have been used as modification reagents. The results obtained are shown in Table 15 (DeVilbiss and Wightman 1986). XPS analysis, utilizing derivatization reactions with fluorine-containing compounds, of functional groups on CFs was also performed by Chan et al. (1991). The results obtained are shown in Table 15. The CFs used were Hercules IM7, Amoco T40, Amoco T65042, and BASF Celion G40800. These CFs are classified as intermediate-modulus grade having a tensile modulus of about 280 GPa. All the fibers were surface treated and contained no sizing. In this case, it is to be noted that though the COOH group is clearly acidic, this is not always true for the hydroxyl group, ROH. The OH group can be acidic if the R group is a phenyl or basic if R is aliphatic. Similarly, the C=O is generally slightly basic, but if there is an alpha hydrogen, such as in CHC=O, this hydrogen atom shows acidic characteristics. Therefore, the assessment of carbon-surface acidity or basicity can only be made tentatively from XPS data, particularly since XPS cannot detect hydrogen atoms (Chan et al., 1991). In such a case, surface acid-base free energy data become useful for deducing whether either one or both of them contribute

to the basic nature or to the acidic nature, as will be shown in Table 16. It has also been pointed out that CFs from various manufacturers have significant differences in their surface acidity (Chun et al., 1992).

1.01.7.3

Surface Free Energy

Chan et al. (1991) made contact-angle measurement to estimate surface free energy on the CFs at the same time as XPS examination of functional groups. In the measurement, methylene iodide was used as the probe liquid for the Lifshitzvan der Waal's interactions; ethylene glycol and formamide were used to probe for the acidbase interactions. The results obtained are shown in Table 16 (Chan et al., 1991). Such surface acidbase free energy data for CFs are applicable for searching a CF more compatible with resin having either acidic, basic, or amphoteric characteristics.

1.01.7.4

Wetting Property

As pointed out by Zielke et al. (1996a), commercial surface-treated carbon fibers are more or less surface-contaminated with adsorbed

Surface Properties
Table 16 Acidbase component of surface free energy for CFs (gLW) ethylene S glycol/CF (gAB7) and formamide/CF (gABg) (all units are in mNm71). SL SL Fiber IM7 T65042 G40800 T40 gLW S 31.4 30.8 31.5 33.2 gAB7 SL 30.1 31.7 24.1 23.8 gAB SL 32.5 27.2 34.8 23.2 1 2 3 4 Rank (amphoteric) (basic) (acidic) (amphoteric)

29

Reproduced by permission of Elsevier Science Ltd. from Chan et al., 1991.

oxidation products. The contaminations, however, can be removed efficiently by extraction with boiling water. Such extraction brings about an increase of the surface concentration of carboxyl groups as a consequence of hydrolysis of anhydrides, which could have been formed by a drying treatment. In scanning tunneling microphotographs on the microand nanoscale, the crevices and pits are more shapely and defined on the washed fiber surfaces, and defects in the form of trenches running longitudinally along the fiber axis are also more pronounced on the extracted fibers which have sharper images (Figure 23). Figure 24 shows the functional groups determined through curve fitting of the O1s peak obtained by the XPS examination of these CFs before and after water washing. For the curve fitting, the following functions were considered: C=O groups (group 1), carbonyl oxygen atoms in esters, amides, anhydrides, and oxygen atoms in hydroxyls or ethers (group 2), the ether oxygen atoms in esters and anhydrides (group 3), and the oxygen atoms in carboxyl groups (group 4). The content of oxygen atoms of group 2 of all fibers varies between 30.4% (fiber (C)) and 52.8% (fiber (E)), and generally surpasses the content of group 1, for which values between 9.1% (fiber (E)) and 32.1% (fiber (F)) were determined. Contents between 5.3% (fiber (F)) and 12.8% (fiber (D)) were obtained for COOH (group 4). Additionally, water was found in most cases, even for dried fibers. The contribution of water was in the range 05.5% in relation to total oxygen. Although XPS is a valuable method for identifying functional groups in a CF surface, it offers an integral picture of a more or less defined surface volume, but not of the net two-dimensional surface. The adhesion of a polymer to the fiber surface is controlled by the surface structure and chemistry of the CFs used. Accordingly, the information of the surface which is most relevant for adhesion can be obtained by contact angle measurement. From such a point of view, Zielke and co-workers have investigated the surface chemistry of CFs by measuring contact angles of the fibers with

aqueous solutions of different pH values (Zielke et al., 1996a, 1996b, 1996c). WSL values were estimated from the contact angles obtained, and WSL/pH value diagrams were obtained for extracted surface-oxidized fibers (unsized Tenax HTA, AS4, T800, and M40) investigated. A schematic presentation of the work of adhesion, WSL, depending on the pH value, is given in Figure 25. WSL is composed of the following fractions (Equations (1) and (2)):
WLW + WAB SL SL WAB = WAB/B + WAB/H SL SL SL (1) (2)

where WSL = work of adhesion, WLW = work SL of adhesion due to Lifshitzvan der Waal's interactions, WAB = work of adhesion due to SL acidbase interactions, and WAB/B = work of SL adhesion due to Bronsted acidbase complexes. LW WSL was estimated from measurements with diiodomethane. Although fibers of different heat-treatment temperatures and crystalline perfection were investigated, those values were in a rather narrow range from 57 to 64 mJ m72. The two kinds of acidbase interactions make contributions to WSL. With all fibers, the main contribution results from hydrogen bonds. At very low pH values they are based on carboxyl, hydroquinone, hydroxyl, phenol, and carbonyl surface groups. With increasing pH value, the carboxyl groups can form Bronsted acidbase complexes. The free energy of complex formation causes an increase of WSL. This can be understood by the following equation:
WAB/B = 7ni DG0 SL i (3)

where ni = number of acidbase complexes i and DG0 = Gibb's free energy of formation of i the acidbase complex i. The height of the step is not large because the contribution of hydrogen bonds formed by carboxyl groups at the lower pH values is simultaneously reduced by the formation of carboxylate anions, and the hydrogen bonds with carboxylate anions are much weaker

30

Polyacrylonitrile (PAN)-based Carbon Fibers

Figure 23 Top views of the surface of CF AS4, unoxidized and unsized, on the micro- and nanoscale. (a) The nonextracted fiber showing less contrast; (b) the extracted fiber with well-contrasted elongated crystallites; (c) a region of the nonextracted fiber with a high degree of surface roughness; and (d) the extracted fiber with well-ordered graphitic regions (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996a 34, 983998).

than those with carboxyl groups. The same interpretation can be given for the second step at high pH values at which Bronsted acidbase complexes are formed between phenolic OH groups and the basic solutions (Zielke et al., 1996c).

1.01.7.5

Reactivity

Generally, the oxidative stability of CFs becomes higher as the HTT or Young's modulus rises. Although CFs are attacked with many kinds of oxidizing reagents such as nitric acid, sulfuric acid, dichromatic acid, kalium-permanganate + acid, etc., even at RT, the temperatures at which the weight loss of CFs starts in oxygen, air, carbon dioxide, and water vapor are in the range 300800 8C. Barr (1995) measured rapid weight loss in air flows at temperatures of 430470 8C, and longterm weight loss in air at 316 8C and 371 8C for CFs having tensile moduli of 207348 GPa. The 60-hour test of 207276 GPa fibers at 430 8C resulted in a weight loss of at least 5%, which is an aid in distinguishing between CFs. The most oxidation-resistant fiber was T-650/50X, which had a modulus of 345 GPa and gave a weight loss of only 1% at 470 8C. Long-term weight loss after 500 h in air at 371 8C for 207276 Gpa fibers ranged from 31% (T-650/42) to 78% (T-300 3K). In contrast, the 345 GPa modulus

fibers were so oxidation resistant at 371 8C that monitoring was continued for 21 000 h (2.4 years). Furthermore, Barr found that the weight losses after 60 h at 430 8C or 450 8C in the rapid test correlated well with the results of long-term aging for 1000 or 2000 h at 316 8C. In the oxidation of carbon materials with gaseous reagents, metallic impurities such as alkali metals and alkaline earth metals accelerate the reaction (McKee, 1981). For example, the dependence of oxidation rates on the sodium content was observed in CFs including

Figure 24 Functional groups determined by XPS showing the relevant oxygen atoms at their corresponding binding energies (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996b, 34, 9991005).

References

31

Figure 25 Model for the use of the adhesion vs. pH value diagram (reproduced by permission of Elsevier Science Ltd. from Carbon, 1996c, 34, 1007 1013).

T-300 (Ismail, 1991). Contamination of CFs with sodium compounds is liable to occur in the process of surface treatment by anodic oxidation using an alkaline electrolite containing a sodium compound, which is one of the commercial treatments. Stark et al. (1994) have investigated the deterioration of an HM CF, Torayca M40, reheated at 3000 8C by means of exposure to atomic oxygen, and found that the Young's modulus and tensile strength decrease with increasing exposure to atomic oxygen, and that the CF surface is severely degraded by atomic oxygen. CFs can be intercalated with various kinds of elements and compounds. The capability of forming intercalation compounds of PANbased CFs, however, is low compared with those of pitch-based CFs and vapor-grown CFs. This is due to the less graphitic structure and circumferential skin structure of the CFs.

1.01.8

REFERENCES

M. R. Alexander and F. R. Jones, Carbon, 1995, 33, 569 580. M. R. Alexander and F. R. Jones, Carbon, 1996, 34, 10931102. S. R. Allen, J. Materials Science, 1987, 22, 853859. Anon., EXPO '70 Document of British Pavilion, Osaka, 1970, 14 (prepared by a British government office, in Japanese). E. M. Asloun, J. B. Donnet, G. Guilpain, M. Nardin and J. Schultz, J. Materials Sci., 1989, 24, 35043510. R. Bacon, J. Applied Physics, 1960, 31, 283290. J. B. Barr, in `Extended Abstracts and Program, 22nd

Biennial Conference on Carbon', San Diego, CA, ACS, University Park, PA, 1995, pp. 1819. Z. Bashir, Carbon, 1991, 29, 10811090. S. C. Bennett and D. J. Johnson, in `Proceedings of the fifth London International Conference on Industrial Carbon and Graphite', 1978, pp. 377386. S. C. Bennett and D. J. Johnson, Carbon, 1979, 17, 2539. S. C. Bennett and D. J. Johnson, J. Materials Sci., 1983, 18, 33373347. G. S. Bhat, F. L. Cook, A. S. Abhiraman and L. H. Peebles, Jr., Carbon, 1990, 28, 377385. D. J. Boll, R. M. Jensen, L. Cordner and W. D. Bascom, J. Composite Materials, 1990, 24, 208219. R. H. Bradley, X. Ling, I. Sutherland and G. Beamson, Carbon, 1994, 32, 185186. W. T. Brydges, D. V. Badami, J. C. Joiner and G. A. Jones, Applied Polymer Symposia, 1969, 9, 255261. D. Chan, M. A. Hozbor, E. Bayramli and R. L. Powell, Carbon, 1991, 29, 10911098. K. J. Chen and R. J. Diefendorf, in `Proceedings of the Fourth International Conference on Composite Materials', Tokyo, eds. T. Hayashi, K. Kawata and S. Umekawa, ISBS Inc., Beaverton, OR, 1982, pp. 97105. Z. F. Chi, T.-W. Chou and G. Y. Shen, J. Materials Science, 1984, 19, 33193324. Courtaulds Ltd., UK Pat. 144 341 (1964). B.-Wa Chun, C. R. Davis, Q. A. He and R. R. Gustafson, Carbon, 1992, 30, 177187. D. Crawford and D. J. Johnson., J. Microscopy, 1971, 94, 5162. G. J. Curtis, J. M. Milne and W. N. Reynolds, Nature, 1968, 220, 10241025. S. J. DeTeresa, Carbon, 1991, 29, 397409. A. Deurbergue and A. Oberlin, Carbon, 1992, 30, 981 987. T. A. DeVilbiss and J. P. Wightman, in `Proceedings of the First International Conference on Composite Interfaces', Cleveland, OH, 1986, pp. 307316. R. J. Diefendorf, in `Symposium on Petroleum-derived Carbon', Philadelphia, American Chemical Society, NJ, 1975, pp. 444445. R. J. Diefendorf and E. Tokarsky, Polymer Engineering and Science, 1975, 15, 150159. N. Dilsiz, N. K. Erinc, E. Bayramli and G. Akovali, Carbon, 1995, 33, 853858. M. G. Dobb, H. Guo, D. J. Johnson and C. R. Park, Carbon, 1995, 33, 15531559. M. G. Dobb, D. J. Johnson and C. R. Park, J. Materials Sci., 1990, 25, 829834. J.-B. Donnet and R. C. Bansal, in `Carbon Fibers', Marcel Dekker, New York, 1984, pp. 1291. J.-B. Donnet and R. C. Bansal, in `International Fiber Science and Technology', Marcel Dekker, New York, 1990, vol. 10, pp. 1470. M. S. Dresselhaus, G. Dresselhaus, K. Sugihara, I. L. Spain and H. A. Goldberg, in `Graphite Fibers and Filaments', Springer Series in Materials Science 5, Springer, New York, 1988, pp. 1382. P. Dunn and B. C. Ennis, J. Applied Polymer Science, 1970, 14, 17951798. P. Ehrburger and J. B. Donnet, in `Handbook of Composites', eds. W. Watt and B. V. Perov, Elsevier, Amsterdam, 1985, vol. 1, pp. 577603. E. Fitzer (ed.), in `Carbon Fibers and Their Composites', Springer, New York, 1985. A. Fourdeux, R. Perret and W. Ruland, in `Proceedings International Carbon Fibres Conference', London, Paper No. 9,1971, pp. 5767 . K. Fujita and Y. Sawada, private communication, 1998. N. Grassie and R. McGuchan, Eur. Polym. J., 1972, 8, 257269. M. Guigon and A. Oberlin, Composites Science and Technology, 1986, 27, 123.

32

Polyacrylonitrile (PAN)-based Carbon Fibers


S. Kumar, D. P. Anderson and A. S. Crasto, J. Materials Sci., 1993, 28, 423439. S. Kumar and T. E. Helminiak, SAMPE J., 1990, 26, 51 61. P. Kwizera, M. S. Dresselhaus, D. R. Uhlmann, J. S. Perkins and C. R. Desper, Carbon, 1982, 20, 387394. G. K. Layden, J. Applied Polymer Science, 1970, 15, 12831285. J. J. Masson and E. Bourgain, J. Materials Sci., 1992, 27, 35273532. R. B. Mathur, J. Mittal, O. P. Bahl and N. K. Sandle, in `Carbon 1992', Preprints of the International Carbon Conference, Essen, The Arbeitskreis Kohlenstoff of the Deutsche Keramische Gesellschft, Koln, 1992, pp. 831 833. R. B. Mathur, J. Mittal, O. P. Bahl and N. K. Sandle, Carbon, 1994, 32, 7177. K. Matsubara, N. Ohba, K. Kawamura, T. Tsuzuku and K. Sugihara, in `Extended Abstracts of the 22nd Biennial Conference on Carbon', San Diego, American Carbon Society, University Park, PA, 1995, pp. 704 705. Y. Matsuhisa, M. Washiyama, T. Hiramatsu, H. Fujino and G. Katagiri, in `Extended Abstracts, 20th Biennial Conference on Carbon', Santa Barbara, CA, American Carbon Society, St. Marys, PA, 1991, pp. 226 227. Y. Matsuhisa, M. Washiyama, T. Hiramatsu, H. Fujino and G. Katagiri, Toray, Private Communication, 1998. D. W. McKee, in `Chemistry and Physics of Carbon', eds. P. L. Walker, Jr. and P. A. Thrower, Marcel Dekker, New York and Basel, 1981, vol. 16, pp. 1118. J. Mittal, O. P. Bahl and R. B. Mathur, Carbon, 1997a, 35, 11961197. J. Mittal, O. P. Bahl, R. B. Mathur and N. K. Sandle, Carbon, 1994, 32, 11331136. J. Mittal, R. B. Mathur and O. P. Bahl, Carbon, 1997b, 35, 17131721. M. Miwa, E. Tsushima and J. Takayasu, J. Applied Polymer Science, 1991, 43, 14671474. R. Moreton, Fibre Science and Technology, 1969, ?vol, 273284. R. Moreton and W. Watt, Carbon, 1974, 12, 543554. K. Morita, H. Miyachi and Y. Kinoshita, in `Preprints of the International Carbon Conference', Baden-Baden, The Arbeitskreis Kohlenstoff of the Deutsche Keramische Gesellschft, Koln, 1972, pp. 303305. K. Morita, H. Miyachi, K. Kobori and I. Matsubara, High TemperaturesHigh Pressures, 1977, 9, 193198. K. Morita, T. Mizushima H. Kitagawa and H. Sakai, US Pat., 3 632 798 (Jan. 4, 1972). K. Morita, Y. Murata, A. Ishitani, K. Murayama, T. Ono and A. Nakajima, Pure Appl. Chem., 1986, 58, 455468. K. Morita, H. Kitagawa, H. Sakai and T. Mizushima, Jpn. Pat., 16 426 (1973). Y. Murakami, K. Nakao, T. Imataki, A. Shindo, K. Honjo and S. Ochiai, in `Composites '86, Proceedings of the Conference on Composites: Recent Advances in Japan and the United States, JapanUS CCM III', The Japanese Society for Composite Materials, Tokyo, 1986, pp. 761766. M. Nakahara and K. Shimizu, J. Materials Sci., 1992, 27, 12071211. Y. Nakayama, F. Soeda and A. Ishitani, Carbon, 1990, 28, 2126. A. Nanjyo, M. Mohri and T. Ishikawa, in `Proceedings of the 39th International SAMPE Symposium', Anaheim, CA, eds. K. Drake, J. Bauer, T. Serajini and P. Cheng, SAMPE, Covina, CA, 1994, pp. 541550. K. Noguchi, K. Murayama and I. Matsubara, in `Preprint of the Fifth Symposium on FRP', Osaka, The Society of Materials Science, Japan, Kyoto (in Japanese), 1976, pp. 14.

M. Guigon, A. Oberlin and G. Desarmot, Fibre Science and Technology, 1984, 20, 177198. A. Gupta and I. R. Harrison, Carbon, 1996, 34, 1427 1445. A. Gupta and I. R. Harrison, Carbon, 1997, 35, 809818. J. Harvey, C. Kozlowski and P. M. A. Sherwood, J. Materials Sci., 1987, 22, 15851596. E. Hayakawa, M. Shioya and A. Takaku, J. Japan Soc. Composite Materials, 1990, 16, 204210. E. Hayakawa, M. Shioya and A. Takaku, J. Japan Soc. Composite Materials, 1994, 20, 187194. J. Heremans, I. Rahim and M. S. Dresselhaus, Physical Review B, 1985, 32, 67426747. T. Hiramatsu, H. Terada and K. Nukada, in `Extended Abstracts Annual Autumn Conference', Chemical Society Japan (in Japanese), 1972, pp. 118119. Y. Hishiyama, Y. Kaburagi and M. Inagaki, in `Chemistry and Physics of Carbon', ed. P. A. Thrower, Marcel Dekker, Inc., New York and Basel, 1991, vol. 23, pp. 1 68. Y. Hishiyama, Y. Kaburagi and A. Yoshida, in `Proceedings of the International Symposium on Carbon', Toyohashi Carbon Society of Japan, Tokyo, 1984, pp. 21. J. W. Hitchin and D. C. Phillips, Fibre Science and Technology, 1979, 12, 217233. K. Honjo and A. Shindo, J. Materials Science, 1986a, 21, 20432048. K. Honjo and A. Shindo, in `Proceedings of the First International Conference on Composite Interfaces', North-Holland, New York, 1986b, pp. 101107. K. Honjo and A. Shindo, Carbon, 1986c, 24, 230234. K. Honjo and A. Shindo, YogyoKyokaiShi, 1986d, 94, 172178. Y. Huang and R. J. Young, Carbon, 1995, 33, 97107. J. D. H. Hughes, Carbon, 1986, 24, 551556. I. M. K. Ismail, Carbon, 1987, 25, 653662. I. M. K. Ismail, Carbon, 1991, 29, 777792. M. K. Jain and A. S. Abhiraman, J. Materials Sci., 1987, 22, 278300. H. Jiang, A. S. Abhiraman and K. Tsui, Carbon, 1993, 31, 887894. D. J. Johnson, in `Chemistry and Physics of Carbon', ed. P. A. Thrower, Marcel Dekker, New York and Basel, 1987, vol. 20, pp. 158. D. J. Johnson and C. N. Tyson, J. Phys. D: Appl. Phys., 1970, 3, 526534. J. W. Johnson, Applied Polymer Symposia, 1969, 9, 229 243. W. Johnson and W. Watt, Nature, 1967, 215, 384386. B. F. Jones, J. Materials Science Letters, 1971, 6, 1225 1227. B. F. Jones and R. G. Duncan, J. Materials Sci., 1971, 6, 289293. C. Jones and E. Sammann, Carbon, 1990a, 28, 509514. C. Jones and E. Sammann, Carbon, 1990b, 28, 515519. W. R. Jones and J. W. Johnson, Carbon, 1971, 9, 645 655. G. Katagiri, H. Ishida and A. Ishitani, Carbon, 1988, 26, 565571. H. A. Katzman, P. M. Adams, T. D. Le and C. S. Hemminger, Carbon, 1994, 32, 379391. T.-H. Ko, H. Y. Ting and C.-H. Lin, J. Applied Polymer Science, 1988, 35, 631640. K. Kogure, G. Sines and J. G. Lavin, Carbon, 1994, 32, 715726. V. V. Kozey, H. Jiang, V. R. Mehta and S. Kumar, J. Mater. Res., 1995, 10, 10441061. C. Kozlowski and P. M. A. Sherwood, J. Chem. Soc., Faraday Trans., 1985, 81, 27452756. C. Kozlowski and P. M. A. Sherwood, Carbon, 1986, 24, 357363. I. Krucinska and T. Stypka, Composites Sci. and Technology, 1991, 41, 112.

References
T. Norita, A. Kitano and K. Noguchi, in `Proceedings of the Fourth JapanUS Conference on Composite Materials', Washington, DC, American Society for Composites, Los Angeles, CA, 1988, pp. 548557. T. Norita, A. Kitano and K. Noguchi, Toray, Private Communication, 1998. M. G. Northolt, L. H. Veldhuizen and H. Jansen, Carbon, 1991, 29, 12671279. Y. Nukushina, J. Matsui and M. Itoh, Sen-i Gakkaishi, 1986, 42, 8489. Y. Nukushina, J. Matsui and M. Itoh, J. Japan Soc. Composite Materials, 1989, 15, 210221. A. Oberlin and M. Guigon, in `Fibre Reinforcements for Composite Materials, Composite Materials Series, 2', ed. A. R. Bunsell, Elsevier, Amsterdam, 1988, pp. 149 210. H. Ogawa and K. Saito, Carbon, 1995, 33, 783788. T. Ohsawa, M. Miwa and M. Kawade, J. Applied Polymer Science, 1990, 39, 17331743. K. Okada, M. Kibayashi, A. Kawaguchi, S. Murakami, H. Iwanaga and Y. Kitano, in `Extended Abstracts of the 22nd Annual Conference, Carbon Society of Japan', Nagasaki, Japan, Carbon Society of Japan, Tokyo (in Japanese), 1995, pp. 150151. C. N. Owston, J. Phys. D: Appl. Phys., 1970, 3, 1615 1626. L. H. Peebles, in `Carbon Fibers', CRC Press, Boca Raton, FL, 1994, pp. 1203. C. U. Pittman, Jr., G.-R. He, B. Wu and S. D. Gardner, Carbon, 1997, 35, 317331. J. M. Prandy and H. T. Hahn, SAMPE Q., 1991, 22, 47 52. A. Proctor and P. M. A. Sherwood, Carbon, 1983, 21, 53 59. M. R'Mili, T. Bouchaour and P. Merle, Composites Science and Technology, 1996, 56, 831834. S. Ochiai and Y. Murakami, J. Materials Science, 1979, 14, 831840. W. N. Reynolds and J. V. Sharp, Carbon, 1974, 12, 103110. W. N. Reynolds and R. Moreton, Phil. Trans. R. Soc. Lond., 1980, A294, 451461. W. Ruland, Applied Polymer Symposia, 1969, 9, 293301. Y. Sawada and A. Shindo, in `Extended Abstracts of the Eighth Annual Conference, Carbon Society of Japan', Kiryu, Japan, Carbon Society of Japan, Tokyo (in Japanese), 1981, pp. 6465. Y. Sawada and A. Shindo, Carbon, 1992, 30, 619629. J. V. Sharp and S. G. Burnay, in `Proceedings of the International Carbon Fibres Conference', London, 1971, pp. 6872. A. Shindo, Report Industrial Research Institute, Osaka, 1961, 316, 152. A. Shindo, in `Abstracts of the Sixth American Carbon Conference', Pittsburgh, PA, held in 1963, Carbon, 1964, 391392. A. Shindo, in `Proceedings of the International Carbon Fibres Conference', London, The Plastics Institute, London, 1971, pp. 1822. A. Shindo, in `Proceedings of the International Conference InterfaceInterphase in Composite Materials', Society of Plastics Engineers-Benelux, Liege, Belgium, 1983, pp. 120. A. Shindo, in `Proceedings of the First International Conference on Composite Interfaces', Cleveland, OH, North-Holland, New York, 1986, pp. 93100. A. Shindo and K. Honjo, in `Proceedings of the Conference on Composites; Recent Advances in Japan and the United States, JapanU.S. CCM-III', Tokyo, North-Holland, New York, 1986, pp. 767774. A. Shindo and Y. Sawada, unpublished results. A. Shindo, Y. Sawada, Y. Nakanishi and K. Honjo, in `Proceedings of the International Carbon Conference', Bordeaux, Groupe France d'Etude de Carbone,
Copyright # 2000 Elsevier Science Ltd. All rights reserved. No part of this publication may be reproduced, stored in any retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or otherwise, without permission in writing from the publishers.

33

Bordeaux, 1984, pp. 144145. I. Shionoya, T. Uchida and K. Nukada, in `Preprints of the International Carbon Conference', Baden-Baden, The Arbeitskreis Kohlenstoff of the Deutsche Keramische Gesellschft, Koln, 1972, pp. 293295. M. Shioya and A. Takaku, Carbon, 1994, 32, 615619. M. Shioya, E. Hayakawa and A. Takaku, J. Materials Science, 1996, 31, 45214532. G. Sines, Z. Yang and B. D. Vickers, Carbon, 1989, 27, 403415. D. H. T. Spencer, M. A. Hooker, A. C. Thomas and B. A. Napier, in `Proceedings of the Third Conference on Industrial Carbon and Graphite', Society of Chemical Industry, London, 1970, pp. 467471. A. K. Stark, L. A. Berglund, M. Tagawa and N. Ohmae, Carbon, 1994, 32, 641644. A. Sumida, K. Ono and Y. Kawazu, in `Proceedings of the 34th International SAMPE Symposium', Reno, NV, eds. G. A. Zakrzewski, D. Mazenko, S. T. Peters and C. D. Dean, SAMPE, Covina, CA, 1989, pp. 2579 2585. T. Takahagi and A. Ishitani, Carbon, 1984, 22, 4346. T. Takahagi and A. Ishitani, Carbon, 1988, 26, 389396. T. Takahagi, I. Shimada, M. Fukuhara, K. Morita and A. Ishitani, J. Polymer Science: Part A, 1986, 24, 3101 3107. A. Takaku, T. Kobayashi, S. Terui, N. Okui and J. Shimizu, Fibre Science and Technology, 1981, 15, 8798. M. Trinquecoste, J. L. Carlier, A. Derre, P. Delhaes and P. Chadeyron, Carbon, 1996, 34, 923929. F. Tuinstra and J. L. Koenig, J. Chemical Physics, 1970, 53, 11261130. F. Tuinstra and J. L. Koenig, J. Composite Materials, 1970, 4, 492499. T. Usami, T. Itoh, H. Ohtani and S. Tsuge, Macromolecules, 1990, 23, 24602465. D. L. Vezie and W. W. Adams, J. Materials Science Letters, 1990, 9, 883887. J. F. Villeneuve and R. Naslain, Composites Science and Technology, 1993b, 49, 191203. J. F. Villeneuve, R. Naslain, R. Fourmeaux and J. Sevely, Composites Science and Technology, 1993a, 49, 89103. X. Wang and D. D. L. Chung, in `Extended Abstracts and Program, 23rd Biennial Conference on Carbon', ACS, Washington, DC, 1997b, pp. 452453. X. J. Wang and D. D. L. Chung, Carbon, 1997a, 35, 706 709. S. B. Warner, L. H. Peebles, Jr., and D. R. Uhlmann, J. Materials Science, 1979, 14, 565572. W. Watt, Proc. Roy. Soc. Lond. Ser. A, 1970, 319, 515. W. Watt, Nature Physical Science, 1972, 236, 1011. W. Watt and W. Johnson, Applied Polymer Symposia, 1969, 9, 215227. W. Watt and B. V. Perov (eds.), in `Strong Fibers Handbook of Composites', Elsevier, New York, 1985, vol. 1, pp. 1575. E. G. Wolff, J. Composite Materials, 1987, 21, 8197. Z. Wu, C. U. Pittman, Jr. and S. D. Gardner, Carbon, 1995, 33, 597605. S. Yamane, A. T. Hiramatsu and T. Higuchi, in `Proceedings of the 32nd International SAMPE Symposium', Anaheim, CA, eds. R. Carson, M. Burg, K. J. Kjoller and F. J. Riel, SAMPE, Covina, CA,1987, pp. 928937. E. Yasuda, Y. Tanabe, H. Machino and A. Takaku, in `Proceedings of the 18th Biennial Conference on Carbon', Worcester, MA, American Carbon Society, St. Marys, PA, 1987, pp. 3031. U. Zielke, K. J. Huttinger and W. P. Hoffman, Carbon, 1996a, 34, 983998. U. Zielke, K. J. Huttinger and W. P. Hoffman, Carbon, 1996b, 34, 9991005. U. Zielke, K. J. Huttinger and W. P. Hoffman, Carbon, 1996c, 34, 10071013.
Comprehensive Composite Materials ISBN (set): 0-08 0429939 Volume 1; (ISBN: 0-080437192); pp. 133

S-ar putea să vă placă și