Vous êtes sur la page 1sur 98

Poly(N-vinylcarbazole):A Selective Review of its

Polymerization, Structure, Properties, and


Electrical Characteristics

R. C. PENWELL,
B. N. GANGULY,
AND T. W. SMITH

Xerox Corporation, Webster, New York 14580

I. Introduction ......................................................... 63
11. Polymerization .... ..................... 64
A. Free-Radical Po

2. Photochemical
E. Copolymerization
F. Polymerization in

III.
IV.
V.
VI.
VII.
VIII.

B. Photoconductivity in Selenium-Sensitized PVK. ...................... .lo9


C. Photoconductivity in Dye-Sensitized PVK . . . . . . . . . 122
D. Photoconductivity of Charge-Transfer Complexes of PVK ............ .127

of N-Vinyl Carbazole ........................ 140


F. Photoconductivity of " ...................
IX. Theoretical Status ................................ 145
A. Simulation Stu
B. Phenomenological Understanding of Photogeneration and
Photoconduction in PVK ......................
1. Photogeneration in PVK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ,146
2. Photoconduction in PVK . . . . . . . . . . . . . . . . . . .
C. Continuous-Time Stochastic Conductivity with Abs
Boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
X. Other Vinyl Carbazoles .......
Acknowledgment ..................................................... 152
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ,153

I. INTRODUCTION

A great deal of attention has been given to poly(N-vinylcarbazole) (PYK)


over the years because of its unusual electrical and photoelectrical properties.

Journal of Polymer Science: Macromolecular Reviews, Vol. 13,63-160 (1978)


@ 1978 by John Wiley & Sons, Inc. OO76-2083/78/0013-0063~01.OO
64 PENWELL, GANGULY, AND SMITH

Earlier work on PVK focused on its dielectric characteristics while more re-
cently its electrophotographic properties have been utilized in commercial
organic photoreceptors. Its unusual electrical behavior, high glass-transition
temperature (227°C) and brittleness, ease of polymerization by a variety of
techniques, and intriguing morphological features and solution properties
combine to make PVK a polymer of considerable current interest.
Because of the more recent applications and commercial use of PVK, this
selective review was written to assimilate presently available information.
The first portion describes much of the work done in the polymerization of
PVK, including polymerization techniques, copolymerization, solid-state and
stereospecific polymerizations. The second portion discusses the solid-state
structure and morphology of PVK as well as its solution, mechanical and
rheological properties, degradative characteristics, and optical properties.
The last portion discusses photoconductivity in PVK from a phenom-
enological viewpoint, including the effects of sensitizers, dyes, charge-transfer
complexes and the photoconductivity of copolymers, binder matrices of N-
vinylcarbazole, and “pretreated” PVK. This experimental discussion is follow-
ed by a summary of the present theoretical understanding and interpretation
of some experimental observations concerning the mechanism of conduction
in organic polymeric photoconductors. A brief discussion of other vinylcarba-
zoles concludes the article.

II. POLYMERIZATION
There are few vinyl monomers which polymerize with the facility of N-
vinylcarbazole (NVK). The rate of its free-radical polymerization is an order
of magnitude faster than that of styrene (Chernobai et al., 1967), and the poly-
merization is quite exothermic; AHDzn= 15.2 & 0.3 kcal/mole (Joshi, 1962).
NVK polymerizes both free radically and cationically. It has been poly-
merized in solution, emulsion (Reppe et a1 , 1935) suspension, bulk (Br. Pat.
739,438, 1955; Ellinger, 1966), and in the crystalline solid state (Restaino
et al., 1957; Galdecki et al., 1967). Its polymerization has been initiated
thermally (Ellinger, 1964; Scott et al., 1963) and photochemically with strong
electron-deficient compounds such as chloranil (Tazuke et al., 1967), as well
as by Ziegler-Natta catalyst systems (Br. Pat. 914,418, 1963; Dall’ Asta and
Casale, 1966; Solomon et al., 1961 ; Heller et al., 1963; Kimura et al., 1970),
and by various metal salts (Tazuke, 1963; Ellinger, 1969).
NVK is a reactive monomer towards radical and cationic polymerization
because of its ability to stabilize electron-deficient centers by resonance in-
volving the nonbonding electron pair on the nitrogen atom in the carbazole
ring. Certain aspects of its polymerization by various catalyst systems are
covered in the following discussion.

A. Free-Radical Polymerization
A quick perusal of the patent literature (Br. Pat. 739,438, 1955; Fikentscher
POLY(N-VINYLCARBAZOLE) 65

et al., 1955; Shimada et al., 1971) suggests considerable commercial int rest
in the free-radical polymerization of NVK, particularly those free-radical
polymerizations carried out in dispersion and suspension. However, there
are only a few published studies on the free-radical polymerization of NVK.
Like many free-radical polymerizations, that of NVK is inhibited by oxy-
gen. Perhaps more important is the effect of certain impurities (anthracene,
phenanthrene, and sulfur) inherent in carbazole derived from coal-tar sources.
The polymerization is retarded and the molecular weight of the polymer
lowered by amounts of anthracene comparable to or somewhat larger than
the initiator concentration, but these effects show saturation at higher an-
thracene concentrations. Phenanthrene is a weak inhibitor which brings about
complete inhibition of polymerization at sufficiently high levels (Ellinger,
1965). It is difficult (although not impossible) to remove the impurities from
coal-tar-derived vinylcarbazole. This can be accomplished by any of several
classical purification schemes including repeated recrystallization, treatment
with Raney Nickel (Davidge, 1959), prepolymerization (Miller and Davidge,
1958), or treatment with maleic anhydride (Bree and Zwarich, 1968). It is
far better to use vinylcarbazole derived from synthetic carbazole because the
latter can be prepared with a high degree of purity (Davidge, 1959).
When pure and oxygen-free, the free-radical polymerization of NVK is
extremely facile. The overall rate constant of polymerizations at 70°C in
cyclohexanone initiated by 0.01M azobisisobutyronitrile (AIBN) is described
by the following equation:

The value of k is 94.5 5~ 2.4 sec-I, as compared to 9.4 f 0.4 sec-l for
styrene (Chernobai et al., 1967). Hughes and North (1966) attempted to de-
termine the individual rate constants for propagation and termination for the
low-temperature solution free-radical polymerization of NVK. The propaga-
tion rate constant expressed in the Arrhenius form is as follows :

The termination rate constant for NVK does not satisfy the Arrhenius rela-
tionship and becomes markedly reduced as the temperature is lowered from
40 to -30°C. This result has been interpreted in terms of the cessation of
backbone rotation in “stiff” macroradicals which markedly retards diffusion-
controlled termination.
Recently, Sutyagin et al. (1974) have reported values of the ratio of the
propagation rate constant k p and the termination rate constant kt for the,
polymerization of NVK in benzene at 70”C, initiated by AIBN. The values
were determined by correlating the rate of polymerization with the average
length of the polymer chain. Since the mechanism of chain termination was
not known, the calculation was carried out for the apparent cases, i.e., for
66 PENWELL, GANGULY, AND SMITH

termination by disproportionation or recombination only; the calculated


values are, respectively,

kp/kt = 11.70 x and 9.20 x 10-2 (liter)l/2 M-1/2 sec-1’2 (3)


In 1972, Jones reported on the benzoyl peroxide (BP0)-initiated polymeri-
zation of NVK. The rate law for the BPO-initiated polymerization is as
follows :

R, = 10-4 [NVK]2[BPO] M (Iiter)-l sec-l (4)

A partial charge-transfer mechanism was proposed to explain the data.


Bevington and Dyball(l975) have carried out end-group studies on PVK
prepared by polymerization at 60°C in benzene using either AIBN or BPO
as initiator. AIBN gave polymers having an average of 1.35 initiator frag-
ments per molecule. BPO gave polymers of unexpectedly low molecular
weight, having close to two initiator fragments per molecule at concentra-
tions of monomer greater than 0.08 M(liter)-l. While this is in keeping with the
mechanism of Jones, other requirements such as direct proportionality be-
tween the molecular weight and the concentration of monomer are not satis-
fied. In fact, the molecular weight is independent of the monomer concen-
tration in the region where the results of end-group analysis are consistent
with the reaction scheme. That a competitive process is operative, in which
simple free-radical polymerization may also be occurring to a small extent,
is indicated by the presence of a small amount of high molecular weight
polymer.

B. Cationic Polymerization
In its free-radical polymerization, NVK is typical of many other vinyl
monomers. In its cationic reactivity, NVK is similar to most other electron-
rich monomers (methylvinyl ether, N-vinylpyrrolidone, etc.). However,
unlike other electron-rich monomers, NVK can be polymerized cationically
even by the weakest initiators. Indeed, it has been polymerized with protonic
acids (Tazuke, 1970). Even the generally accepted concept that basic solvents
preclude cationic polymerization cannot be applied to NVK (Solomon et
aI., 1964; Tazuke et al., 1965).
Boweyer et al. (1971) have studied the cationic polymerization of NVK
by tropylium hexachloroantimonate and tropylium perchlorate. Polymeri-
zation initiated by both catalysts showed identical features; there was a
virtually instantaneous reaction followed by very rapid propagation. Yields
of polymer were invariably quantitative and the evidence pointed to absence
of termination during the kinetic lifetimes. Molecular weights of PVK from
the perchlorate initiation were slightly lower than those from the hexachloro-
antimonate catalysis and did not show the same temperature variation. From
both catalysts, however, the molecular weight data showed clear evidence
POLY(N-VINYLCARBAZOLE) 67

of a transfer reaction characteristic of homogeneous cationic systems.


Although reactions between tropylium salts and NVK were instantaneous
under most conditions, at temperatures below -50°C the two reagents reacted
to form only a pink-colored solution from which polymer could not be precip-
itated on addition of methanol. However, if the solutions were allowed to
warm to around - 30" C before addition of methanol, polymer formed rapidly.
It seems likely, therefore, that initiation involved preequilibrium formation
of a charge-transfer (CT) complex which collapses rapidly above - 50"C,
yielding a propagating cation.
Reaction rates were measured by an adiabatic calorimetric technique and
half-lives were of the order of 1-2 sec. Catalyst concentrations were suffi-
ciently low (< lO-5M) for essentially complete dissociation into free ions, as
indicated by ion-pair dissociation constants. The rate constant for propaga-
tion by free cations in methylene chloride at 0°C was estimated as -3 x 10-5
M-1 sec-1, approximately five orders of magnitude greater than for the cor-
responding free-radical polymerization.
Ledwith and Sherrington (1972) have also reported the only example of a
propagating dication by reaction of NVK with the stable cation-radical
tris-(p-bromophenyl) aminium hexachloroantimonate. The polymerization
involves the reaction of the stable cation radical with NVK to form carbazyl
cation radical which either dimerizes directly (in a manner analogous to the
Na/naphthalene/styrene anionic system) or in two steps via oxidation of the
radical addition product with a second molecule of (BrC6H&N (t) SbClsc-).
Although this system with the propagating dication is extremely interest-
ing and important, a question remains as to why relatively large amounts
of catalyst must be used and why only a molecular weight of - 10,000 can
be achieved. Ledwith (1975) has suggested that this may be caused by com-
petitive cyclodimerization.
There are numerous other reports of the cationic polymerization of NVK,
utilizing a variety of acidic reagents (S-hildknecht, 1949; Biswas, 1967; Fr.
Pat. 1,397,538, 1965; Miyama et al., 1967; Neth. Pat. Appl. 6,612,244,
1967; Bawn et al., 1964; Ledwith, 1967; Cassiers and Hart, 1971 ;Biswas and
Kamannarayana, 1975; Hernandez and Bandini, 1973). Among the most
recent studies are those by Pielichowski (1973) who described the polymeri-
zation of NVK initiated by hydrogen halides, HF, HCI, HBr and HI; by
Biswas (1970, 1973, 1974) who has studied the polymerization of NVK by
chlorides and oxychlorides of some group V elements and by Turchi et
al. (1971) who have examined the polymerization of NVK by triethyloxonium
hexafluorophosphate.

C. Electroinitiated Cationic Polymerization


There have been a few reports in the literature of the electroinitiated
polymerization of NVK, one of the most recent being a report by Kikuchi
and Fukuda (1974) of the electroinitiated polymerization of NVK in acetic
anhydride and acetonitrile solutions containing lithium perchlorate. These
68 PENWELL, GANGULY, AND SMITH

workers suggested that acyl groups or hydrogen cations which were produced
by electrolysis of the solvent initiated the cationic polymerization of the
monomer.
Kikuchi and Ueyama (1974) have subsequently reported the electrochemi-
cal polymerization of NVK by rectangular waveform current (RWC). The
electrolysis of a solution of NVK and lithium perchlorate in acetic anhydride
or acetonitrile by 5:1 RWC resulted in the formation of NVK in the anode
compartment of a divided cell. No polymer was obtained when various
solvents other than acetic anhydride or acetonitrile were used. The intrinsic
viscosities of the polymers formed by RWC were higher and their colors
were lighter than those polymers formed in the original work by direct
current. Ultraviolet and ir examination showed that more ester and acyl
groups existed in the polymer formed in acetic anhydride by RWC than
by DC; the structure of the polymers formed in acetonitrile were similar to
those of polymers produced by standard cationic initiators.
The first report was that by Breitenbach and Srna (1962) resulting from
the passage of current through solutions of NVK in nitrobenzene containing
silver perchlorate. Phillips et al. (1972, 1973) and Davies et al. (1973)
have recently reported that the electroinitiated reaction of NVK in zinc
bromide/acetone solutions can result in either linear polymer or cyclodimer,
truns-1, 2-dicarbazylcyclobutane (TCB). The product depends only on the
current density used in the electrolytic cell. It should be noted that as early
as 1964, Breitenbach reported that the electroinitiation of NVK with HgII
cyanide in acetonitrile gave a cyclic dimer.
Ledwith (1972) and Crellin and Ledwith (1975) have unequivocally im-
plicated the NVK cation radical in both the metal-catalyzed and photochemi-
cal cyclodimerization of NVK. In the following section of this paper
ample evidence of the cationic propagation of NVK(+) will be presented.
It is concluded that in many cases of electroinitiated polymerization of NVK,
the initiating species is the NVK cation radical.

D. Polymerization by Electron-Deficient Species


In the last decade perhaps the most active area of research in the polymeri-
zation of NVK has centered on its polymerization by electron-deficient spe-
cies, including electron acceptors [chloranil, tetracyanoquinodimethane
(TCNQ), tetracyanoethylene (TCNE), tetranitromethane, benzoquinone,
etc.]; haloalkanes (CX4, CHX3; X = C1, Br, or I); electron-poor alkenes
(maleic anhydride, acrylonitrile), and metal salts (NaAuCl4, AgC104, AgBFT,
NaAuBr4). These polymerizations have been initiated both thermally and
photochemically. Whereas there is some understanding of the mechanisms
of the photochemical reactions, the situation with respect to the thermal
reactions is ambiguous at best.

I . Thermal Interactions
In the early 1960s, Scott et al. (1963) and Ellinger (1964, 1965) almost
POLY(N-VINYLCARBAZOLE) 69

simultaneously reported the polymerization of NVK by organic electron


acceptors. These included quinones such as chloranil ; aromatic polynitro-
compounds such as tetranitrobenzene and 1-chloro-2,4-dinitrobenzene;
electron-poor alkenes such as maleic anhydride, trichlorethylene, and
tetrachlorethylene.
Pac and Plesch (1967) carried out a kinetic analysis of the thermal poly-
merization of NVK by tetranitromethane in nitrobenzene. The results indi-
cated that the polymerization was cationic in nature. They postulated a slow
ionogenic dissociation of the CT complex between monomer and catalyst
with chain transfer by monomer maintaining an almost constant very low
concentration of rapidly growing chains. A purely cationic propagation was
supported by the fact that similar kinetics were observed in transfer experi-
ments with the catalysts tetranitromethane, chloranil and HCl.
Early reports of the thermal polymerization of NVK by CT complexes
were disputed, first by Meyers (1966) and later by Natsuume et al. (1970),
who showed independently that purified chloranil did not thermally initiate
polymerization of NVK in benzene or toluene. The active cationic initiator
was a hydrolysis product of chloranil (3,5,6-trichloro-2-hydroxy-l ,4-benzo-
quinone), a strong protonic acid. Similarly, the indications are, despite con-
siderable controversy (Scott and Labes, 1963; Breitenbach and Olaj, 1964;
Chapino and Hardy, 1962; Biswas and Glosal, 1966; Scott et al., 1964),
that the various thermal polymerizations of NVK in halocarbon solvents
are cationic in nature and catalyzed by acidic species.
So far there has been little clear evidence for the formation of NVK
cation radical from thermally induced reaction of CT complexes (with the
SCHEME I
NVK + TCNE [CTC] : I[NVK”] [TCNE”]

-
-
CN CN

sLowl - HCN

FA ST
I >
70 PENWELL, GANGULY, AND SMITH

possible exception of the polymerization of NVK by TCNE) (Bawn et al.,


1971 ; Nakamura et al., 1970). NVK reacts with TCNE to form a 2 + 2-
adduct, the butadiene derivative (1,1,2-tricyanobutadienylcarbazole)(Will-
iams et al., 1962) and PVK. CT spectra are thought to represent common
reaction intermediates, and the cationic polymerization of NVK could be
initiated either by TCNE or by the cycloadduct, as depicted in Scheme I
on p. 69 (Hyde and Ledwith, 1974).

2. Photochemical Interactions
With respect to donor-acceptor photopolymerization in general and the
photochemical polymerization of NVK by electron acceptors in particular,
a recent review by Tazuke (1973)discusses the present state of understanding
very well. Inasmuch as only certain aspects related to NVK will be described
here, the reader is encouraged to read Tazuke’s entire review.
The photoinduced cationic polymerization of NVK in nitrobenzene was
the first definite example of initiating ionic polymerization of a vinyl com-
pound by CT excitation (Tazuke et al., 1967). Since this work, various photo-
polymerizations of NVK using both organic and inorganic acceptors have
been investigated and a general kinetic relationship of the form

R, oc Zo1’2 (5)

was found (Asai et al., 1971) where R, is the rate of propagation and I0 is
the intensity of incident radiation. Propagation is always cationic.
To explain the rate data, it is required that the termination mechanism
involve the recombination of solvated anion radicals with the growing carbon-
ium ions.

P(+)+ A(:) -+ termination


If recombination occurs within the solvent cage, the kinetic expression R,
cc does not hold. An alternate explanation is to assume the coupling
of primary active species.

D ( t ) + A(:)
\
-
slow -
DA
pzn

If solvation of the excited CT complex (DA)* or exciplex, D*A or DA*, is


inefficient and if only a small fraction of ion radicals function as initiators
for cationic polymerization, the rates of polymerization would again vary
with assuming a linear termination process for the propagating species.
Specific examples which have been studied are NVK-nitrobenzene (Tazuke
et al., 1967; Tagawa et al., 1974), NVK-sodium tetrachloroaurate-nitro-
benzene (Tazuke et al., 1968;Asai et al., 1974), NVK-carbon tetrabromide-
ethylene chloride with and without acridine orange (sensitizer) (Olaj et al.,
POLY(N-VINYLCARBAZOLE) 71

1971; Kauffmann et al., 1973), and NVK-silver perchlorate-aromatic hydro-


carbon (Asai et al., 1975; Takeda et al., 1975; Tazuke, 1973).
In the polymerization of NVK initiated by sodium chloroaurate (NaAuC14.
2Hz0) in nitrobenzene at 30" C, the rate of polymerization is proportional to
[AuIII] [NVK]. When reducing agents are added to the system, the rate of
propagation in the dark increases. The rate of propagation is relatively unaf-
fected by addition of water and N-ethylcarbazole, but polymerization is com-
pletely inhibited in the presence of ammonia. Free-radical scavengers [OZ
and 2,2-diphenyl-l-picrylhydrazyl(DPPH)] act neither as inhibitors nor
retarders. Kinetic treatments based on the assumption that the active ini-
tiating species is A@, produced by reduction of Au'II by NVK and other
reducing agents, explain the experimental results very well (Tazuke, 1968).
To avoid the complexity in light absorption presented by photoreactions
in nitrobenzene, anhydrous gold halides, ( ~ - C ~ H ~ ) ~ N ( + ) A U X
=~Cl,
( - )Br),
(X
which are more soluble in common organic solvents were used in a study of
the initiation mechanisms of photocationic polymerization (Asai and Ta-
zuke, 1973). Polymerizations were carried out under 436-nm irradiation in
ethylene dichloride, tetrahydrofuran, nitrobenzene, and acetophenone. All
polymerizations were characterized by a cationic propagation step with a
sigmoidal time-conversion curve. In acetophenone, the bromide and the
chloride complexes showed nearly equal reactivity and the light-absorbing
species was the gold salt itself, although some complex formation with NVK
was observed for the chloride complex. After complete decomposition of the
gold complex in the early stage of the photopolymerization, thermal polymeri-
zation in the dark continued to complete conversion. It was concluded that
the initiating species was produced by the reaction of NVK with a transient
AuII or Br atom as photochemical redox products of AuBr;.
In more recent reports, the cationic polymerization of NVK by Ag(1)
perchlorate and tetrafluoroborate in various aromatic solvents has been de-
tailed (Asai et al., 1975; Takeda et al., 1975). Under irradiation at 365 nm
at 30°C, the polymerization rates decreased in the following order : nitroben-
zene > benzene > toluene >> p-xylene -0. The effective absorption species
leading to initiation is NVK itself. The proposed initiation mechanism
entails electron transfer from excited NVK* to solvent, which produces a
cation radical, NVK(f), which in turn reacts with AgCI04 to produce the
initiating species.
The photochemical reaction of NVK with an organic acceptor molecule
most often leads to formation of the NVK cation radical, which can undergo
other reactions, including free-radical or cationic propagation. The fact
that cationic polymerization of NVK is much faster than the radical process
leads to the observation of cationic propagation.
The photopolymerization of the NVK-acrylonitrile (ACN) system affords
an opportunity to study this duality (Ellinger, 1964; Tazuke and Okamura,
1968a, 1968b). It should also be remembered that with the formation of
the NVK(+),the possibility of forming cyclodimer always exists.
Photopolymerization of the NVK-ACN system irradiated by a high-
12 PENWELL, GANGULY, AND SMITH

pressure mercury lamp in a glass ampoule (A > 300 nm) produces a mixture
of copolymer and homopolymer (Tazuke and Okamura, 1968a).In the presence
of NH3, the formation of PVK is entirely suppressed, whereas the addition
of DPPH enhances the relative amount of PVK to copolymer. These results
demonstrate the simultaneous occurrence of radical and cationic polymeri-
zation under the influence of photoirradiation.
Because absorption spectroscopy of the NVK-ACN system (Tazuke, 1970)
does not provide any evidence of CT interaction in the ground state and be-
cause fluorescence of NVK is quenched by ACN, this photopolymerization
is likely to be exemplary of the situation in which the donor, NVK, is first
excited and subsequently reacts with the acceptor, ACN, to form an exciplex.
In addition, the absorption by ACN at wavelengths longer than 300 nm is
negligible, and thus the photoenergy is mostly absorbed by NVK (see
Scheme 11).

SCHEME I1

NVK hV * NVK* * NVK* ACN

radical propagation +
c o p lymer

c a t i o n i c propagation

I
homopolymcr

As additional proof for the intermediacy of NVKW in photolysis of NVK-


acceptor systems, the photoionization of NVK in the presence of various
acceptors has been studied by means of flash photolysis (Shirota et al., 1972;
Yamamoto et al., 1973), the first example being the NVK-chloranil system
(Kimura et al., 1970) in which NVK? chloranil(:) were observed spectro-
scopically.
Dissociative reduction of the carbon-halogen bond seems to be a common
trend of CT reactions involving halogenated acceptors and the production
of H-X generally in the form of the ammonium salt has been described (Ste-
venson and Coppinger, 1962). With respect to photopolymerization systems,
reaction products such as HCl are capable of initiating polymerization if the
monomer is susceptible to cationic polymerization. Photopolymerization of
NVK-haloalkane systems would thus fall into the category of initiation by
stable reaction products via CT interaction. A typical example is the NVK-
POLY(N-VINYLCARBAZOLE) 73

CBr4 system (Olaj et al., 1971). The NVK forms a CT complex with CBr4
in ethylene dichloride; irradiation at 436 nm excites the CT complex ex-
clusively and induces cationic polymerization of the monomer (Tazuke,
1973).
Kauffmann et al. (1973) have recently reported on the photochemical
dimerization of NVK in the presence of CBr4 and oxygen. In methanol
solution the cationic polymerization is suppressed and no high-molecular-
weight polymer is formed. However, in the presence of small amounts of
oxygen (2 x lO-3M) photodimerization occurs leading to the cyclodimer,
trans-l,2-dicarbazole-9-yl-cyclobutane, and the open-chain dimeric dime-
thoxy derivative, 1,4-dicarbazole-9-y1-1’,4’-dimethoxy-n-butane.

E. Copolymerization of NVK
The copolymerization of NVK has been studied in connection with CT
interactions of NVK with electron-poor alkenes (Ellinger, 1965; North and
Whitelock, 1968). In this regard, claims have been made of the preparation
of alternating copolymers (Chapiro and Mankowski, 1973; Shirota et al.,
1972; Yoshimura, 1973). As long ago as 1949, the copolymerization of
NVK was being studied when Alfrey and Kapur (1949) reported the following
reactivity ratios r for the polymerization of styrene (Sty) and methylmeth-
acrylate (MeMA) with NVK :

M1 Mz
- rl
- r2
Sty NVK 5.5 0512
MeMA NVK 2.0 0.20

The Polymer Handbook (Brandrup and Immergut, 1974) lists reactivity


ratios for the polymerization of NVK with a variety of vinyl monomers,
including ally1 halides (Alfrey et al., 1952a,b), styrene and styrene derivatives,
acrylates and methacrylates (Alfrey and Kapur 1949; Alfrey and Magel,
1961; Hurt, 1961), vinyl esters (Ushakov and Nikolaev, 1956a-c), and
vinylidine chloride (Hurt, 1961).
The first documented report of the alternating copolymerization of NVK
was made by Shirota et al. in 1972 when these workers described the thermal
AIBN-initiated copolymerization of NVK and fumaronitrile (FN). The
copolymer composition was found to be always constant, irrespective of
widely varied monomer feed ratios, and to be almost 1 :1 (NVK/FN). The
maximum initial copolymerization rate occurred at a monomer feed ratio
of 7:3 (NVK:FN). In absence of AIBN, cationic homopolymerization of
NVK was initiated by FN.
These workers have since reported on the alternating free-radical copo-
lymerization of NVK with diethyl fumarate (DEF) (Yoshimura et al., 1973)
diethyl- and dimethylmaleate (Yoshimura et al., 1975), as well as the photo-
induced polymerization of NVK with FN and DEF (Shirota et al., 1973;
Tada et al., 1973). The photopolymerization was initiated by photogener-
74 PENWELL, GANGULY, AND SMITH

ated cation radicals of NVK and the relative basicity of the solvent deter-
mines whether one gets radical copolymerization, cationic homopolymeriza-
tion of NVK, or cyclodimerization of the NVK cation radical.
Limburg and Seanor (1975) have prepared intramolecular charge-transfer
complex copolymers of N-vinylphthalimide (NVP) and NVK via the free-
radical (AIBN) polymerization of NVP and NVK in benzene. Isostructural
modifications of these copolymers have also been carried out. Hydrazinolysis
of the NVP residues yields vinylamine/NVK copolymers, which have been
reacted with tetrachlorophthalic anhydride to give tetrachloro-NVP/NVK
copolymers and with 2,4-dinitrofluorobenzene and 1-fluor&nitrobenzene
to give copolymers of N-viny1-2,4-dinitroaniline/NVK and N-vinyl-4-
nitroaniline/NVK, respectively (see Scheme 111).

NO2
Scheme I11
Turner (1975) has found that I-phenyl-l,2,4-triazoline-3,5-dionespon-
taneously copolymerizes with NVK to give low molecular weight alternating
copolymers :

oi.;Lo \

8' -Ab,)- &


NVK
LN-N

0
0- CH -CH&-

Stolka and Turner (to be published) have discovered that diethylazodi-


carboxylate can be free-radically copolymerized with NVK. The process is
normally accompanied by a parallel homopolymerization of NVK. However,
in basic solvents which suppress the cationic polymerization of NVK, the
only product obtained is an alternating copolymer :
POLY(N-VINYLCARBAZOLE) 75

Block and graft polymerization offers the possibility of getting better me-
chanical properties in NVK-containing polymers. Accordingly, the synthesis
of free-radical “graft” copolymers of polybutadiene and NVK has been re-
ported (Seanor, 1975). NVK was “grafted” to polybutadiene in solution and
in emulsion. Although these “graft” copolymers were more flexible and ex-
hibited better adhesion to aluminum than PVK, it was concluded that rubber
reinforcement of PVK did not lead to such improvements in mechanical pro-
perties as those observed in styrene/butadiene systems. In terms of photocon-
ductivity, it was found that the “graft” copolymers were slightly less efficient
than PVK.
Stolka (1976) has employed a novel synthetic method for the prepara-
tion of block copolymers of NVK. The method, studied on NVK and
N-dodecylmethacrylate as model systems, is a two-step radical polymeriza-
tion initiated by sequential initiators such as di[ 1,3-dimethyl-3-(t-butyl-
peroxy)butyl]peroxydicarbonate. The product is a multiblock copolymer
(AB)n formed along with the homopolymer of NVK. The homopolymer can
be extracted from the block copolymer; however, it has been found that in
the situation where polymerization occurs with precipitation of polymer,
block copolymers virtually free of homopolymers can be obtained. The latter
situation results in long-living “trapped” radicals in the solid phase of the
newly formed polymer. A series of multiblock copolymers of NVK and laur-
ylmethacrylate with molecular weights up to 3 x 106 and containing 50-90 %
were synthesized. The average molecular weight of each block can be con-
trolled by reaction conditions in each step of sequential polymerization.

F. Polymerization in the Crystalline Solid State


The solid-state polymerization of NVK may be carried out by the action
of high-energy radiation (Galdecki et al., 1967; Kroh and Pekala, 1964, 1966;
Ayscough et al., 1968; Mankowski, 1973), cationic catalysts (Okamura et
al., 1965, 1966; Meyers and Christman, 1968; Nishii et al., 1964, 1966; Ma-
tsuda et al., 1967; Tusuji et al., 1964; Solomon et al., 1965), and free-radical
initiators (Matsuda et al., 1968; Boros-Tyevi, 1969). Investigations in the
field of solid-state radiation-induced polymerization show that the structure
of monomer crystals directly affects the course of the reaction (Barkalov et
al., 1964; Amagi and Chapiro, 1962) as well as the properties of polymers
(Banford et al., 1960; Okamura et al., 1962). According to Okamura, the li-
near spatial distribution of polymer in a solid monomer requires the poly-
merization to proceed along a crystallographic axis. The effect of monomer
76 PENWELL, GANGULY, AND SMITH

structure on the course of radiation polymerization of NVK has been studied


(Galdecki et al., 1967; Kroh et al., 1966). Microscopic examination of mono-
crystals of NVK irradiated with x rays revealed the presence of fibrous
polymer aggregates formed in the crystal during polymerization. The oriented
distribution of these aggregates suggested that the polymer growth was
determined by the crystallographic axis of NVK. However, the polymer
obtained was not crystalline (Kroh and Pekala, 1964). It has been suggested
that in the case of NVK, the destruction of crystallinity may be caused by an
increase in the volume of the macromolecule.
A recent paper (Hardy et al., 1973) has considered the effect of phase
relations of monomerpolymer binary systems on solid-state polymerization.
+
Based on detailed experimental studies of the acenaphthylene polyacenaph-
thylene and the NVK + PVK systems by polaroptical, DTA, and x-ray dif-
fraction methods, it was concluded that in a given polymerization at a critical
composition (Mcr),three phases are in equilibrium at a characteristic tempera-
ture : the viscous polymer-monomer liquid, the highly elastic polymer-mono-
mer system with composition Mcr,and the crystalline monomer. At polymer
contents lower than Mcr, the system consists of one or two phases. One of
the two phases is the crystalline monomer and the other i s a monomer-poly-
mer “solution” which, as the temperature changes, may be liquid, highly
elastic, or glassy. The phase can be described approximately by the funda-
mental relation for plasticized polymers as long as the polymer concentration
is higher than Mcr. Below Mcr an approximate description is possible,
considering the nonideal solubilities of the crystalline materials.

G. Ziegler-Natta Polymerization of NVK


There has been considerable interest in the polymerization of NVK by
Ziegler-Natta catalyst systems because of the possibility of forming stereo-
regular crystalline (crystallizable) polymers. Among the reported systems
which have been used are EtAlClz, EtzAlCl, Ti(OPr)zClz, Ti(OAc)zC12,
Fe(OAc)Clz, Sn(0Ac)zClz (Br. Pat. 914, 418, 1973; Dall’Asta and Casale,
1966), BuLi-TiCl4, Al(iso-Bu)a-TiCls, and AlEt3TiC14 (Solomon et al., 1961;
Heller et al., 1963; Kimura et al., 1970).
Contradictory results have been reported on the crystallinity of PVK pre-
pared by Ziegler-Natta catalyst systems, Solomon et al. (1961) and others
(Br. Pat. 914,418, 1963; Kimura et al., 1970)reporting the polymer to be crys-
talline; Heller and coworkers (1963) finding it noncrystalline. Kimura et al.
(1970) observed that under a variety of conditions (radical, cationic, or Zie-
gler-Natta) PVK powder precipitated from solution is a material of low crys-
tallinity. The x-ray diffraction curves for all polymer samples were essential-
ly the same. These workers further reported that the degree of crystallinity
increased in all polymer samples on annealing. It is apparent from this and
other studies that, aside from the question of whether or not PVK is crystal-
line, there are strong indications that the x-ray structure of the polymer is
POLY(N-VINYLCARBAZOLE) 77

independent of the method of polymerization. That is to say, the bulky carba-


zole group (not the method of polymerization) is determining the structure
of the polymer. The question of polymer structure will be discussed in more
detail in Sec. IV.

III. THERMAL STABILITY STUDIES


Since PVK is a vinyl derivative with a high T,, requires high processing
temperatures (>230°C), and sees environments in photoelectric applications
such as uv light, oxygen, and corona discharge, the molecule is expected to
degrade. Very little work has been done in the past on the thermal and radia-
tive degradation, oxidation, and ozonization of PVK. Other vinyl polymers
have been studied extensively (Tsuji, 1973; Grassie and Weir, 1965a-d;
Guillet, 1972; Grassie, 1972). Recently work has appeared which discusses
the kinetics of thermal degradation of PVK (Barrales-Rienda et al., 1975)
and its pyrolysis (Chu and Stolka, 1975).
Decomposition of polymers can occur via a direct or indirect process.
Direct decomposition occurs when chromophores or irregular bonds are
contained in the main chain (Tsuji, 1973). Indirect decomposition includes
energy-transfer processes and radical reactions following absorption by impu-
rities. The possibilities for energy transfer are discussed elsewhere. Whether
by direct or indirect decomposition, the formation of a free radical can lead
to one or more of the following reactions (Tsuji, 1973);
ROO. Inert

RH -+ R. +ROO*
0 2
r --* ROOH
ROOH
+ R---+ RO- + HzO + ROO.
(A) (C) RH
@) L R O -+ .OH
L+HzO + R.

k R O H + R.
Step A is associated with the initiation reaction and is influenced by the
presence of radical initiators, oxidation products, and sensitizers. Step B is
associated with propagation and is influenced by the presence and diffusion
of oxygen. Steps C are causes of branching and cross-linking in the polymer,
and are affected by the presence of easily removed hydrogens, uv, heat, or
transition metals (Tsuji, 1973).
Oxidation of the pendant groups is also possible. In vinyl polymers it is
initiated by an attack on the tertiary hydrogen and, in some cases, by an at-
tack directly on the pendant group (Stokla, 1975). Generally, the aromatic
groups oxidize to form quinones, which can then undergo further reactions
to form alcohols (Stolka, 1975).
Thermal, oxidative, and environmental degradation tests have been per-
78 PENWELL, GANGULY, AND SMITH

formed on several commercial and fractionated PVK materials as a function


of ambience, temperature, and time (Bergfjord and Penwell, 1975). Gel-per-
meation chromatography was used to monitor the degradation. Purified,
low molecular weight samples were the most resistive to degradation in air
and argon at 180,280, and 330°C. Purified high molecular weight PVK per-
formed better than an "as-received" commercial material which contained
approximately 5 % impurities. The impurities evidently initiate and help
propagate degradative reactions leading to a discolored cross-linked matrix.
The differences observed in air and argon are due to the presence of oxygen
which, according to the suggested reactions, increases the propagation re-
action and leads to branching and cross-linking.
In the absence of an abundance of free radicals, and considering the bulk
of the carbazole group, stripping of the monomer under vacuum might be
expected. Chu and Stokla (1975) recently completed a study on the pyrolysis
of PVK. Using a purified BASF PVK material ( a n = 340,000), the pyrolysis
was carried out in vucuo and the low molecular weight products were distilled.
PVK pyrolyzed in V ~ C U Oat temperatures ranging from 230-410°C for 18 h
remained soluble. A 0.2% weight loss occurred at 260°C in 18 h. Within the
same time period, several percent weight was lost above 300°C. At tempera-
tures below 350"C, the weight loss is due to the stripping of monomer. At
the highest temperatures, the distilled product contained small amounts of
low molecular weight polymer. At 410°C, the polymer was completely pyro-
lyzed in 18 h. At 350"C, a tenfold decrease in a n occurred with only 25%
volatilization. Their conclusion was that the cleaved macromolecule of PVK
has a limited tendency to eliminate monomer from the active ends. Some
monomer is released followed by termination via disproportionation. PVK
in a nitrogen atmosphere in the range 2 3 0 4 0°C becomes cross-linked, in-
soluble, and colored after 18 h. They conclude that since vacuum-pyrolyzed
PVK remains soluble and the only primary degradation product is monomer,
the formation of carbazole and cross-linking are secondary processes.
A study by Barrales-Rienda et al. (1975) was done in nitrogen and it has
been observed that this can greatly affect the degradation process (Chu and
Stolka, 1975; Bergfjord and Penwell, 1975). It appears that the kinetic para-
meters and activation energies they derive only describe the distillation of
PVK in nitrogen under their test conditions. The mechanism(s) involved in
the thermal degradation of PVK have not yet been firmly established nor is
the mechanism of carbazole formation known (Chu and Stolka, 1975; Bar-
rales-Rienda, 1975).
Qualitative environmental effects such as light, nuclear radiation, aging,
and corona discharge have been considered (Fritz, 1964; Lanker, 1972). De-
gradation and cross-linking generally result in poorer mechanical properties
and may affect the electrical characteristics of PVK (Tsuji, 1973; Jellineck,
1967; Lenz, 1967). A decrease in tensile properties and increased embrittle-
ment may result.
79

IV. SOLID-STATE STRUCTURE

The discussion of ordering in amorphous PVK and the possibility of solu-


tion or bulk crystallization begin with a consideration of the stereoregularity
of the polymer. The presence or absence of stereoregularity refers to the con-
figuration of the atoms or monomer units in the polymer backbone which
result from primary bond formation. This is separate from the conformation
of the polymer, which is determined by the rotation of atoms around single
bonds (Billmeyer, 1971). The generally accepted description of stereoregular
vinyl polymers comes from a consideration of the fully extended planar zigzag
conformation. If all the substituent groups on the vinyl polymer lie above or
below the plane of the main chain, it is isotactic. An alternating placement is
the syndiotactic configuration while a random sequence is atactic. As pre-
viously discussed, the synthesis technique selected may result in a predeter-
mined configuration.
The presence of regularity of molecular structure enhances the possibility
of crystallization. Another important consideration in determining the po-
tential for crystallization is the bulkiness of the vinyl substituent group. Its
structure and size affect the conformation of the chain and hence the packing
ability (Billmeyer, 1971).
There is presently much discussion about whether PVK will polymerize
stereoregularly(Br. Pat 914,418., 1963; Heller et al., 1963; Kimuraet al., 1970;
Crystal, 1971;Williams, 1970; Griffiths, 1975; Solomon et al., 1961). The bulki-
ness of the carbazole group would seem to disfavor isotactic formation even
with the use of stereospecific Ziegler-Natta catalysts. Solomon et al. (1961)
say they obtained a highly isotactic PVK polymer using organometallic cata-
lysts and obtained spherulitic structures. Heller et al. (1963), using similar
Ziegler-Natta-type catalysts, observed no indication of crystallization. They
did concede that the large bulk of the carbazole side group might prevent
crystallization and hence they were not entirely sure they did not have stereo-
regular PVK. Others have also claimed to have synthesized stereoregular
PVK (Br. Pat 914,418., 1963; Kimura et al., 1970).
Nuclear magnetic resonance is perhaps the most effective tool at our dispo-
sal in studying the tacticity of macromolecules. Yoshimoto et al. (1969) were
the first to assign the protons of PVK:

-CH -2H T = 8.5


-CH2 -1H 5 = 6.4 and 1.3
Ar - H -1H z = 5.0
Ar - H s multiplet 5 = 3.0

The difference in the chemical shift for the two methine protons was attributed
to restricted internal rotation of the PVK molecules in solution. The extremely
-
large upfield shift of one aromatic proton to z 5 was explained by the shield-
ing effect of neighboring carbazyl groups.
80 PENWELL, GANGULY, AND SMITH

Williams (1970) has studied nmr spectra of PVK in several solvents at 60


and 220 MHz and at a series of temperatures (see Figs. 1 and 2). The tempera-
ture dependence of the nmr spectrum of PVK as well as the nonuniform
shielding of the aromatic protons suggests that there is hindered rotation within
the polymer segments with unequal population of conformers. Spectra of
PVK in 1,4-dichlorobenzene at 60 MHz do not show any temperature depen-
dence of the relative intensities of the main components of the methine proton
doublet at temperatures up to 175" C . The lack of temperature dependence of
the components of the doublet as well as their upfield shift from the methine
proton septet in n-isopropylcarbazole suggests that this multiplicity is due to
differences in shielding of the methine proton lines in the various stereochemi-
cal sequences and is not due simply to restricted internal rotation.
References to oriented PVK first appeared a number of years ago, mostly
related to the processing of PVK (Beck and Dorrer, 1940; Busse and Lam-
bert, 1952; Busse et al., 1948; Davidge, 1957; Cornish, 1963). Generally, the
improved mechanical properties of PVK extruded through annular dies, noz-
zles, and slits provided the evidence for the orientation of PVK. Very little

r
Fig. 1. 60-MHz nmr spectrum of PVK in pyridine-d5 at a series of temperatures. Re-
printed with permission from the author, Williams (1970). Copyright by the American
Chemical Society.
81

C
I35.C

Fig. 2. 220-MHz nmr spectrum of PVK in pyridine-ds at a series of temperatures. Re-


printed with permission from the author, Williams (1970). Copyright by the American
Chemical Society.

characterization or morphological work was carried out on the PVK mater-


ials. Kimura et al. (1970) discuss the crystalline microstructure of PVK pre-
pared by radical, cationic, and Ziegler-Natta catalyzed polymerizations. They
detected no difference in crystallinity among the differently polymerized sam-
ples. By thermally annealing the sample at 280°C under pressure for 30
min and subsequently stretching five times at 20C220" C , they obtained their
most crystalline material. From their x-ray diffraction work, they concluded
that the PVK molecules were rigid, rodlike, and packed approximately par-
allel in a pseudohexagonal array. They obtained a lateral spacing of 10.7 h;
and interpreted it as {lOiO} diffraction of the pseudohexagonal lattice. As-
suming this to be correct, the nearest chain-to-chain distance is 12.6 h;. No
longitudinal order was detected.
Examination of models suggested the possibility of an isotactic 3/1 helix
with TGTG conformation or a syndiotactic 2/1 helix consisting of TGTG'.
Crystal (1971), using x-ray and electron diffraction and electron microscopy
82 PENWELL, GANGULY, AND SMITH

and optical microscopy, studied the morphology of a commercial PVK


having = 2.5 x lo5. He obtained “pseudospherulitic” structures at
temperatures above 285”C with shear-induced crystallization observed at
temperatures as low as 230”C. The latter sample consisted of fibrous bundles
approximately 250 in diameter lying parallel to the shear direction. He
also was able to obtain interesting morphologies crystallizing from dilute
solution. Below 115°C he obtained a rod-type morphology with a structure

Fig. 3. Schematic diagram of a cut through PVK rod like molecules arranged in a hexa-
gonal array, indicating key crystallographic orientations. Reprinted with permission from
the author, Crystal (1971). Copyright by the American Chemical Society.

I I 1
CRYSTALLIN: PEAK
10.78A

I AMORPHOUS “HALO=

8.2 10 2023 Jo

Fig. 4. X-ray diffraction traces for both crystalline and amorphous PVK. Reprinted with
Reprinted with permission from the author, crystal (1971). Copyright by the American
Chemical Society.
POLY(N-VINYLCARBAZOLE) 83

similar to that reported by Kimura et al. (1970). As seen in Figure 3, a1 = a2 =


a3 = 12.3 A
as confirmed by x-ray and electron diffraction data. X-ray
diffractometer traces for both a crystalline and an amorphous PVK sample
are shown in Figure 4. The crystalline peak at 28 = 8.2" represents an in-
terplanar spacing of 10.78 A.
The amorphous sample is characterized by an
amorphous region at 28 = 20-25" and a weaker reflection in the vicinity of
the crystalline peak. Crystal (1971) also observed that PVK polymerized by
several techniques crystallizes in the same manner. His observations therefore
suggest that all the polymerizations be considered to lead to stereoregular
polymers. Furthermore, the assignment of PVK by Kimura et al. (1970) and
A
himself to a hexagonal array with c = 7.4 requires that the PVK be highly
isotactic and form a rod with a 3/1 helix conformation. Griffiths (1975) has
recently studied the structure of PVK in which discrepancies with prior work
appear. He utilized a high molecular weight fraction (a,
= 2.3 x 106, a,/
d, = 1.25) in his electron microscopy and electron and x-ray diffraction
measurements. This fraction was solvent cast onto evaporated carbon and
aluminum substrates and was crystallized from 200-330" C under nitrogen
with bulk and free-standing films. Nucleation and subsequent crystal growth
resulted when T > 270"C. Bulk and free-standing films gave regularly stacked
lamellar structures and no spherulites. X-ray scattering curves from amor-
phous films gave diffuse peaks at 28 equal to 7.9 and 21.5". Crystalline films
gave a single sharp peak at 28 equal to 8.39"(10.39 A)
and increased
resolution at 14" (6.3 A).
His x-ray curves are seen in Figure 5.

In
k
I I I 1
z _ - _ - - -AMORPHOUS FILM (GANOOLFI)
3

:
-
- CRYSTALLINE FILM (GANDOLFI)
t
m
- - - -CRYSTALLINE FILM
(DIFFRACTOYETER 1
a
a
I -
F
c_
In
z -
W
c
z
0 -
z
a
w -
b
b
a ,
4
v
u l -

F
a . -- ,'

'Ik
a
X

0 8.
SCATTERING ANGLE - 2 8

Fig. 5. Curve similar to Fig. 4. Reprinted with permission from John Wiley and Sons,
Inc. and the author, Griffiths (1975).
84 PENWELL, GANGULY, AND SMITH

Selected area transmission electron diffraction from single lamellae of crys-


tals grown on carbon gave a sharp single-crystalp?ttern. With the hexagonal
symmetry existing about the chain axis, the 10.39 A reflection was interpreted
as the {lolo} reflection of a pseudohexagonal lattice with a = 12.00 A. It
was also concluded from the size of the lamellae in the chain-axis direction
that chain folding must occur.
Considering an isotactic vinyl polymer chain with a 3/1 screw axis and the
trigonal symmetry suggested by hexagonal packing, the diffuse reflection at
14"(6.3 A) is interpreted as the {OOOl}reflection of the crystal lattice. Griffiths
(1975) then calculates the c-axis dimension using the area of the unit cell in
the {OOOl} plane and the volume of the unit cell calculated from the mole-
cular weight/monomer unit and the 100 % crystalline density. The relation
between measured densities and percent crystallinity determined from x-ray
scattering data was extrapolated to obtain the 100% crystalline density of
PVK. Details of the determination may be found in Griffiths' (1975) original
work. His results are presented in Figure 6 as the percent crystallinity and
calculated c-axis/monomer unit vs. density relations. Values of 1.193 g/cc for
100% crystallinity and c axis equal to 2.16/monomer unit or 6.47 A for a
3/1 isotactic chain are obtained. The latter value agrees fairly well with the

2-20*
2.19

2.141 I

Fig. 6. Percent crystallinity and chain axis period/monomer unit as functions of density.
Reprinted with permissionfrom John Wiley and Sons, Inc. and the author, Griffiths (1975).
POLY(N-VINYLCARBAZOLE) 85

6.3 A diffuse peak in the crystalline x-ray curve. A density difference of about
1 % between the amorphous and paracrystalline states suggests an amorphous
state with considerable order. Griffiths' (1 975) idealized molecular packing
model for PVK is illustrated in Figure 7.

T-
6.476

Fig. 7. Structure model for crystalline PVK. Reprinted with permission from John Wiley
and Sons, Inc. and the author, Griffiths (1975).

Griffiths' (1975) results eliminate the possibility of a 2/1 syndioactic chain


and confirm the presence of a 3/1 isotactic chain. Crystal's (1971) main chain-
axis periodicity, when compared to the results of Griffiths' (1975), and the
published amorphous density, gives a density approximately 20 % too low.
Stein and Su (1975) have recently obtained evidence of crystallization in
the bulk from their rheo-optical studies of PVK. Their samples were prepared
by casting films 3-10 mils thick from a 10% PVK/tetrahydrofuran solution.
The final drying conditions were 160"C , then 250" C in a vacuum oven over-
night. No residual solvent analysis on the films was made. Their morphologi-
86 PENWELL, GANGULY, AND SMITH

cal evidence was obtained using birefringence, ir dichroism, small-angle light


scattering, and wide-angle x-ray techniques.
The observed samples were subjected to a variety of tensile deformation
histories as a function of temperature and time (Su, 1975) and birefringence
measurements were used to study structural changes. By definition, birefrin-
gence ( A ) is equal to the difference in refractive indices in different directions
and therefore will indicate morphological changes in uniaxial extension.

where nl and n2 are the refractive indices parallel and perpendicular to the
stretching directions, l o is the wavelength of light in vacuum, d is the sample
thickness, and R = d/ll - d/lz ( l 1 , l z are the wavelengths of light parallel
and perpendicular to the stretching direction). The order of retardation, R ,
was measured with a Babinet compensator. Stein and Su (1975) observed an
increase in the birefringence at 240°C and for extension ratios 2 4. For the
system described, this is characteristic of ordering the molecules, and crys-
tallization was thought to have occurred. Similar observations were made at
other combinations of temperature and extension ratios.
Stein and Su (1975) use Eq. (7) to calculate the intrinsic birefringence L&
of the oriented crystalline component in PVK. From the Denbigh bond po-
larizabilities, Jcwas found to be 0.46 :

A = n 1 - n2 = -
217 (ii-2 + 2)2 (Pl - P z )
9 n- (7)

where ri is the mean value of the refractive index and pl and pz are the bond
polarizabilities parallel and perpendicular to the stretch direction.
Stein (1958) has described the birefringence of a multiphase system as

where 41 and Ai denote the volume fraction and birefringence in the ith phase
and A f is the form birefringence representing the contribution from any aniso-
tropic boundaries. For semicrystalline PVK, Su (1975) wrote Eq. (3) as

where Ae and do are the birefringence contributions from the crystalline and
amorphous components, respectively. He redefines Ae and A a in terms of
intrinsic birefringence d” and average orientation function f, and Eq. (9) be-
comes

He uses the following linear relation between birefringence and stress for
amorphous polymers:
POLY(N-VINYLCARBAZOLE) 87

Aa = C.O (1 1)
where C is the proportionality constant (called the stress optical coefficient)
and o is the stress. Neglecting the form birefringence, he writes Eq. (10) as

A =&
j& + C o(l - qL) (12)
Rearranging and with fc = 1,

Stein and Su (1975) have used Eq. (13) to determine the percent crystal-
linity occurring during their stress relaxation studies. An example of their
results is shown in Figure 8 where the percent crystallinity is given as a func-
tion of time for various extension ratios at 285°C. Their wide-angle x-ray
measurements produced results similar to those discussed by Griffiths (1975).
The ir dichroism measurements were essentially in agreement with their bire-
fringence measurements (Su, 1975).

T = 285'C 0 =2.0
A 1 =3.0
X =40
A A = 5.0
.Ot
40 -

$ 30-
0
X

20 -

LO -

0 _, 0 " - T I

Time (min 1

Fig. 8. The variation of crystallinity with time at 285OC for extension ratios of 2.0 to 5.0.
Reprinted with the permission of the author, Su (1975).

The low-angle light-scattering results of Stein and Su (1975) also suggest


an appreciable amount of orientation exists. At 275" C and a stretch ratio of
two, they observed a four-leaf clover in the Hv scattering pattern (electric
vector of polarizer is parallel to the stretching direction and analyzer is per-
pendicular), indicative of spherulitic structure. At stretch ratios of three and
88 PENWELL, GANGULY, AND SMITH

greater, an x pattern was observed which is characteristic of fibrillar scatter-


ing. Similar results were observed at other temperatures.
Griffiths (1976) has recently studied the crystallization of PVK as a function
of molecular weight. The propensity to crystallize decreases with decreasing
molecular weight and is zero below a number-average molecular weight (an)
of 46,000. He feels the 3/1 helix existing in solution must nucleate the crystal
growth with the chain-end free energy controlling the stability of the PVK
folded-chain nuclei.
Recent melt extrusion of a commercial PVK (a, = 1.5 x lo6, gw/an
5.5) resulted in what appeared to be a highly uniaxially oriented film (Pen-
well and Prest, 1976). Initial density, x-ray, and ir dichroism measurements
suggest the possible presence of a one-dimensional hexagonal array with
interchain spacing similar to that measured by Griffiths (1975).
Some of the earlier work on PVK may be criticized for using uncharacter-
ized commercial samples, for not monitoring retained solvent in solution-cast
films, or for not analyzing end products for possible degradation. However,
the more recent rheo-optical studies and x-ray and electron diffraction work
on characterized PVK provide strong evidence for a high degree of order in
PVK (Kimura et al., 1970; Crystal, 1971; Griffiths, 1975, 1976; Su, 1975).
Although there appear to be discrepancies in recent work on the degradation
of PVK, there is evidence which suggests that degradation may occur at an-
nealing temperatures above 300" C (Chu and Stolka, 1975; Bergfjord and Pe-
nwell, 1975;Barrales-Rienda et al., 1975). The relatively small density change
of less than 1 % in the transition from the amorphous to crystalline state is
representative of paracrystalline behavior (Griffiths, 1975, 1976). The degree
of order in PVK proposed in the models will be further clarified when the
tacticity and microstructure are fully resolved by nmr.
The orientation of PVK by drying solvent-cast films in the presence of a
uniform electric or magnetic field has recently been reported (Nagao and Her-
mann, 1974). Birefringence measurements were similar in magnitude to those
of stressed amorphous polymers. No evidence of crystallinity using back-
reflected x rays was obtained.
Dielectric and relaxation time studies have resulted in tentative assignments
for the a,p,y, and 6 relaxation regions observed in PVK (Froix et al., 1975;
Pochan et al., 1975). The a process has been associated with the onset of large-
scale translational displacement of the polymer chains associated with the
glass-transition region. A Tgof 225" C agrees well with calorimetric date (Berg-
fjord and Penwell, 1975)7. From a consideration of the activation energies
of the p and 6 processes and comparative data in the literature, the /Itransi-
tion of 80" C was attributed to a carbazole wagging motion and the 6 transition
at - 160"C to a relaxation associated with a carbazole-localized backbone
motion (Pochan et al., 1975). The y relaxation at - 60" Cis due to ring libration
of the carbazole group about its symmetry axis. Froix et al. (1975) found the
y relaxation active only in the presence of oxygen. A frequency-temperature
map of the observed transitions from the dielectric and nmr data is shown in
Figure 9 and the corresponding activation energies are noted.
89

9- I I I I I I 1
o DIELECTRIC (J. Pochon)
8- 4 A + NMR
-
I \
7 -
I \ -
LOG y
I \
6-
\ -
I \
I \ Y E,-B.Of0.8
5- I\
4 - "\ \ T Bl lEo'19f3
M o l e

-
2-
a E0%77KcOVMolr
I 1 I I I I 1 1 1 I

Fig. 9. Correlation frequency uC vs. l / T f o r PVK dielectric and nmr data. Reprinted with
permission from the American Institute of Physics and the authors, Froix et al. (1975).

V. SOLUTION PROPERTIES

Further effects of the bulky carbazole substituent on the conformation of


PVK can be observed by looking at its solution properties. Several studies
of solution properties have been made in which the parameters and relations
discussed below give the most interesting information about PVK (Klopffer,
1969a; Ueberreiter and Springer, 1963; Ueberreiter and Bruns, 1964; Hughes
and North, 1964; Naghizadeh and Springer, 1967; Van Krevlin and Hoftyzer,
1967; North and Phillips, 1968; Kuwahara et al., 1969; Sitaramaiah and
Jacobs, 1970). With the proper conditions and solvent, much can be learned
about the conformation of PVK in solution and bulk.
Using the classical random-flight method, the root-mean-square end-to-
end distance of a dissolved freely jointed chain is given by

where x links of length I are joined in a linear sequence (Flory, 1953). Because
of restrictions to free rotation and long-range interactions resulting from
excluded volume (two chain atoms cannot occupy the same volume), real
polymer chains deviate from this ideal behavior. In the absence of long-range
interactions, a condition obtained by using a suitable solvent and temperature,
90 PENWELL, GANGULY, AND SMITH

the chain dimensions are called the unperturbed dimensions and can be repre-
sented by 3.A characteristic ratio can now be defined as

which is used as a quantitative measure of polymeric chain flexibility. To


account for long-range interactions, the unperturbed dimension 5,
based on
a statistical analysis, must be multiplied by an expansion factor a. This
accounts for the mutual exclusion of segments and gives
-
( r 2 ) 1 / 2 z ,(;?,)1/2 (16)
where LY depends on the solvent used and the temperature. If a solvent and
temperature where LY = 1 are found, the unperturbed dimensions of the
polymer can be obtained (Flory, 1953). These conditions are called the 8 tem-
perature and 19solvent. A brief explanation of how the condition is found and
what additional information it can give about PVK follows. In early theories
concerned with frictional properties of polymers in solution, the intrinsic
viscosity [q]was found to depend on the volume of the molecule in solution
divided by a molecular weight (Flory, 1953). Consideringthe linear dimension
of a chain given in Eq. (14), the expression for [q]is

where @ may be considered constant (Flory, 1953). Substituting Eq. (16),

[q] = @ ( $ / M ) 3 / 2 M1%3 (1 8)
From Eq. (14) and remembering (q)1/2
is an unperturbed dimesnion, Eq.
(1 8) can be written as

Under 8 conditions, a = 1, giving

This can be expressed in a more general form as

Based on hydrodynamic theory, Einstein obtained a relation for a dilute sus-


pension of hard spheres which results in a = 0 in Eq. (21) (Flory, 1953). In
the Debye theory for a free-draining molecule (which considers the work done
on a molecule, represented by a series of beads on an infinitesimal string,
placed in a shearing fluid), Eq. (21) has a = 1 (Flory, 1953). In consideringthe
POLY(N-VINYLCARBAZOLE) 91

thermodynamics of a swollen polymer molecule where CY >> 1, it can be shown


that (Y = K'M1/10(Flory, 1953). Considering Eq. (19), this gives a = 0.8
corresponding to a partially draining coil. Generally, the molecular weight
exponent in Eq. (21) lies between 0.5-0.8. If the coefficient a is determined
for PVK in any solvent, something is then known about the conformation
in solution and whether the solvent is good or bad. If a 0 solvent is used,
the ratio (rT/M)1/2can be determined, thus allowing a calculation of the char-
acteristic ratio c in Eq. (1 5). As mentioned, this is an indication of the steric
hindrance in the polymer. The unperturbed dimension of the molecule ( 2 ) 1 / 2
can be calculated, giving an indication of the conformation of the polymer in
bulk. Modifications have been made to the solution equations presented to
account for the non-Gaussian behavior with excluded volume, but these are
not discussed here (Sitaramaiah and Jacobs, 1970).
Naghizadeh and Springer (1967) determined the 8 temperature to be 39 f
2" C in toluene. Kuwahara et al. (1 969) observed a B temperature of 37 f 2"C
in toluene in their work with PVK fractions. The latter obtained values of
K = 7.62 x and a = 0.50 (0 condition) for the constants in Eq. (21).
Determinations of K and a in other solvents and at other temperatures are
seen in Table I. The work of Sitaramaiah and Jacobs (1970) suggests that at
25" C benzene is a poor solvent and tetrachloroethane is a very good solvent.
As pointed out by Kuwahara et al. (1969), the most reliable unperturbed
dimensions are obtained by using a 8 solvent. Having determined K in Eq.

TABLE I
K and a values in [ ] = KMa for PVK

Solvent Temperature
2
("C) K(xl0 ) a
Benzene (Sitaramaiah and Jacobs, 1970) 25 3.05 0.58

Cyclohexanone ( I b i d . ) 25 2.00 0.61

Tetrahydrofuran ( I b i d . ) 25 1.44 0.65

Chlorofonn ( I b i d . ) 25 1.36 0.67

Tetrachloroethane ( I b i d . ) 25 1.29 0.68

Benzene (Ueberreiter and Springer, 1963) 25 3.35 0.58

Toluene (Kuwahara e t a l , 1969) 37 7.62 0.50

For light-scattering measurements, the refractive index for wavelengths of 436 and 546
nm have been reported for several solvents and temperatures (Ueberreiter and Bruns, 1964;
Sitaramaiah and Jacobs, 1970). The second virial coefficient Az has been determined by os-
motic pressure and light-scattering measurements for various molecular weights and
solvents at different temperatures (Ueberreiter and Springer, 1963 ;Naghizadeh and Spring-
er, 1967; Kuwahara et al., 1969; Sitaramaiah and Jacobs, 1970). Measurements of r 2 and
the mean-square radius of gyration have also been reported under various conditions
using intrinsic viscosity and light-scatteringdata by the same coworkers.
92 PENWELL, GANGULY, AND SMITH

(21), the ratio ($/M)3/2 is now determined since @ in Eq. (1 8) is known from
theoretical calculations. Kuwahara et al. (1969) obtained a value of 6.33 x
10-9 for ( 2 / M ) 1 ' 2 with @O = 2.87 x (nondraining flexible coiled poly-
mers). This value agrees quite well with ( s / M ) 1 / 2= 6.19 x 10-gobtained by
Sitaramaiah and Jacobs (1970) using a different PVK-solvent system. The
characteristic ratio cr for PVK in Eq. (15) is 2.85 as determined by Kuwahara
et al. (1969) and 2.82 as determined by Sitaramaiah and Jacobs (1970). In
comparison to cr values for other polymers, PVK is quite high and indicates
the bulky carbazole pendant groups have a large steric repulsion (Brandrup
and Immergut, 1974; Flory, 1953; Kuwahara et al., 1969; Sitaramaiah and
Jacobs, 1970). The volume of the pendant group, Vc,in PVK correlates well
with the data Privalko (1973) used to establish the relation Ve = 4.4 cr3.5. This
expression is valid for polymers with a helical conformation in the crystalline
state.
For the solvents listed in Table I, the values of the exponent in Eq. (21)
suggest that PVK is a partially free-draining coil in solution. The hypochro-
mism of PVK in chloroform has recently been measured by Tsuchihashi et
al. (1975). They have measured the diminution of the uv absorption intensities
of the carbazolyl groups at various molecular weights relative to those for
PVK with molecular weight 1200. The degree of diminution of the bands is
caused by specific interactions between the pendant chromophores. The
interactions are a result of the hindered internal rotation of the carbazolyl
groups around the main chain due to their bulkiness. They have attempted
to correlate the hypochromism with the structure in solution. The intensities
diminish when the molecular weight is increased from 1200 to approximately
10,000, after which they are constant. This suggests regular orientation of
the pendant carbazole rings in the chloroform solution. They feel there is a
rigid helical structure at lower molecular weights, possibly interconnected by
random and more flexible coils at higher molecular weights. They do not
state or do not know what the limiting size of the rodlike structure is. Nu-
clear magnetic resonance information they present suggests the rigid local
structure of PVK only begins to form at a molecular weight of 1200. Their
data are consistent with the work of Dev et al. (1970) concerning dielectric
and viscoelastic relaxation measurements of dilute solutions of PVK. The
dielectric relaxation studies of North and Phillips (1968) also confirm that
PVK molecules form very stiff structures in solution. The behavior in solution
proposed by Tsuchihashi et al. (1975) appears consistent with other available
information; i.e., PVK gives the appearance of a rigid rod in solution, but
has characteristics associated with a partially free-draining coil at higher
molecular weights.
Iyer et al. (1975) incorporated up to 20 % comonomer of methyl acrylate
and styrene into the backbone of PVK. Their acoustical studies showed no
effect on the intrinsic rate of segmental motion in the hindered PVK chain.
They concluded that the barriers to segmental rotation have their origin in
long-range dipole-dipole interactions between carbazole groups as well as in
POLY(N-VINYLCARBAZOLE) 93

the short-range steric effects. The rotating unit contains several monomer
residues and so is little affected by a chemical defect.

VI. MECHANICAL AND RHEOLOGICAL PROPERTIES


Not much has been published concerning the mechanical properties of
PVK. A large share of the materials that have been measured has been
poorly characterized with respect to molecular weight distribution, impurities,
and solvent content. The glass transition of purified (impurity levels <
0.01 %) PVK is 225 2°C (Griffiths, 1969; Bergfjord and Penwell, 1975).
Therefore, it behaves as a glass in the temperature range of normal interest
in response to mechanical treatment. The glass transition temperature T,
decreases with molecular weight according to the relation T, = Tgm-
K / a , where Tpm 227°C and K = 2.27 x 105 (Bergfjord and Penwell,
1975). This relation is shown in Figure 10. The limiting value of T, at high
molecular weights is TBwand a n is the numger-average molecular weight.
Impurities will also reduce T,. A commercial PVK obtained from BASF
(aw = 1.55 x lo6, a w / n ? n 5.5) containing 4.5% carbazole plus traces
of other impurities has a T, of 200°C (Bergfjord and Penwell, 1975). The

490 -
0
400 -
z
0
I-
470-
z
a
a
t-
v) 460-
v)
a
J
0
450-

440 -

Fig. 10. The glass-transition of PVK as a function of number-average molecular weight.


94 PENWELL, GANGULY, AND SMITH

glass-transition temperature can be lowered with the addition of low molecular


weight compounds compatible with PVK. The Tgof purified PVK decreased
=: 26°C by the addition of 25% of a low molecular weight PVK fraction
having aw = 4500 (Bergfjord and Penwell, 1975).
A list of mechanical properties of PVK is given in Table 11. Most of
these measurements were reported a number of years ago when the materials
were not sufficiently characterized (Beck and Dorrer, 1940; Busse and
Lambert, 1952; Busse et al., 1948; Davidge, 1957; Cornish, 1963). The
excellent thermal stability of the mechanical properties is offset by the
extreme brittleness of PVK. The reported low elongation ratios of 0.32%
(Cornish, 1963) for PVK limit its mechanical usefulness.
Although many glassy polymers undergo a yielding phenomenon, most
fail because of crazing and/or fracture because the yielding stress exceeds
the critical stress for fracture (Legrand, 1969). The criterion for the brittle-
ductile transition is based on the magnitude of the stress necessary for the
growth of the flaw versus the stress to cause yielding. The stress for growth
of the flaw is a, = JEy/C, where E is Young’s modulus, y is the surface energy,
and C is the size of the flaw. If a, > oY,ductile failure results, whereas brittle
failure results if ay > ag (Legrand, 1969). The process for brittle fracture is
a dilational process. Therefore, annealing near Tg,plasticization, orientation,
compressive prestressing, and hydrostatic pressure are all potential ways to
transform a glassy polymer from brittle to ductile behavior (Legrand, 1969;
Rabinowitz and Beardmort, 1972; Golden et al., 1967; Mininni et al.,
1973). In several glassy systems, there is evidence suggesting that mechanical
properties may increase with a molecular weight increase even after the
glass-transition temperature and other material parameters do not change
(Foden et al., 1972).
The possibility of plasticizing PVK to improve its processing characteris-
tics and mechanical properties was first mentioned a number of years ago
(Busse and Lambert, 1952; Busse et al., 1948; Cornish, 1963; Orinik, 1948).
The list of plasticizers discussed may be found in the original papers. Most
reduce the viscosity, thus permitting processing at lower temperatures.
Chloroform is suggested as a plasticizer which reduces interfacial adhesion
and permits easy stripping of PVK film (Orinik, 1948). The use of fillers
and foaming agents has also been discussed in earlier work (Beck and Dorrer,
1940; Stastny and Gerlich, 1955; Stastny and Buchholz, 1956; Dumont and
Reinhardt, 1957; Ellinger, 1966).
The improvement of mechanical properties by orientation was first dis-
cussed a number of years ago. Beck and Dorrer (1940) discussed orienting
PVK by extrusion through nozzles and slits. Since then, a number of other
workers have mentioned the improvement in mechanical performance by the
extrusion of fibers of PVK and extrusion through annular dies (Busse and
Lambert, 1952; Busse et al., 1948; Davidge, 1957; Cornish, 1963). Recently,
Prest (1974) oriented thin films of PVK by stretching near the glass transition,
obtaining a highly flexible film. Melt extrusion of PVK through a thin slit
POLY(N-VINYLCARBAZOLE) 95

TABLE I1
Mechanical Properties of PVK (Cornish, 1963)

P r o prt y "Polectron" "Luvican"

5
W u l u s (lb/in2 5.8 4.7 x 10
2
Tensile s t r e n g t h ( l b / i n ) 1800 2050

18,000-20,000
(oriented f i b e r s )

Elongation (a) 0.32

Impact s t r e n g t h ( f t - l b / i n ) 0.19 2.2

Impact bending stress (cm/kg,cm) 10-15


2
Flexural strength (Ib/in 1500-8000 5500-6900

(increasing with
f i b e r content)

4500-5500 (-2O'C
to +80°C)
2
Hardness (kg/m ) 14

(1) Rockwell R 125

(2) Ilackwell 113

Compressive s t r e n g t h ( l b / i n 2 ) 4800
2
Shear s t r e n g t h ( l b / i n ) 3500

S p e c i f i c h e a t (cal/°C/gm) 0.3

Thermal c o n d u c t i v i t y (cal/cm/sec°C) 6.0

C o e f f i c i e n t of l i n e a r expansion
(/"C) 45-55 x 57 x
(2O-10O0C)

Heat d i s t o r t i o n temperature (ASTM l0O-l5O0C


method)
Heat r e s i s t a n c e (Martens) 150°C
(Vicat) 19ooc

Flow temperature 27OoC

Permissible continuous temperature 12oOc

Specific gravity 1.2 1.2

W f r a c t i v e index 1.69 a t 20°c

Water absorption 2
0.1% 12 mg/100 c m /week

Polectron is a trademark of General Aniline & Film Corp.


Luvican is a trademark of Badische Anilin-und-Soda-Fabnk.
96 PENWELL, GANGULY, AND SMITH

die resulted in highly uniaxially oriented film which is highly flexible in the
direction of extrusion. X-ray, ir, density, and birefringence measurements
indicate the presence of a high degree of order in this material (Penwell and
Prest, 1976).
Since PVK has such a high glass-transition temperature, the melt-pro-
cessing temperature is expected to be a great deal higher than 230°C. In
earlier work on PVK, softening points of just over 200°C were mentioned
(Davidge, 1957). Undoubtedly, the PVK was of lower molecular weight or
contained plasticizers in the form of impurities. Rheological data on charac-
terized PVK materials are very limited. One reason is that the onset of de-
gradation appears quickly at nominal melt temperatures. Erhardt and
Richards (1975) measured the zero shear viscosity no of commercial BASF
(a, = 1.3 x 106) and Desoto (aw = 2.5 x lo5) PVK at 285°C. The no
values were 2.0 x lo7 and 1.5 x lo5 P respectively. Dynamic viscoelas-
tic measurements in the shear mode using a Rheovibron were made on a
series of eleven PVK fractions ranging in molecular weight from a, =
5450 to Bw= 1.3 x lo6. Frequencies of 3.5, 11, 35, and 110 Hz were used
at temperatures from 176" to 300" C. From these measurements, the average
molecular weight between entanglement points Mc was estimated to be
35,000. Their Tgmolecular weight constant K, estimated from the T g - Bde-
pendence of modulus, was 2.1 x 105, which is in good agreement with the
value mentioned earlier.
As another means of measuring orientation in uniaxially stretched films,
Erhardt and Richards (1975) determined the strain optical coefficient for
PVK. The birefringence/unit strain was found to be -5.5 k 1 x 10-2 in the
range 210-235" C and was relatively independent of temperature and stress
level. Stress optical coefficients have also been reported (Su, 1975).
Stein and Su (1975), from their stress and birefringence relaxation meas-
urements, obtained master curves covering several decades of reduced time.
The data followed the time-temperature superposition principle, allow-
ing them to evaluate the constants in the WLF equation (Ferry, 1971). For
a reference temperature of 220" C and from the stress relaxation measure-
ments, values of C1 = 11.4 and CZ = 226.0 were calculated (Su, 1975).

VII. OPTICAL PROPERTIES OF PVK

Useful insights into the complex spectroscopic properties of organic


polymers are often gained by studies of appropriate model compounds.
The uv spectra of PVK, solution and (film), exhibit maxima a t 29,100
(29,100); 33,950 (33,800); 38,250 (38,150); and -43,000 cm-l (see Fig. 11).
By comparison with the solution spectrum of N-isopropylcarbazole (NIPC)
it can be seen that neither the spectral position of the transitions nor their
absorption intensities are altered. This demonstrates that the carbazole
chromophores of the polymer act as independent units in the light absorption
process, both in solution and in the solid state. This is not the case for the
POLY(N-VINYLCARBAZOLE) 97

lo5
&

10‘

,I l i / I I ,I I
1”lyl

lo2

10’
55 50 45 40 35 30 25 20.W’
cm-’
Fig. 11. The UV absorption spectra of PVK dissolved in methylene chloride (1) and as a
film (2). Reprinted with permission from the American Institute of Physics and the author,
Klopffer (1969a).

emission process in PVK where both excimer and monomer fluorescence


have been observed (Klopffer, 1969a). In the case of PVK, the next step in
complexity, above that of the carbazyl group itself, is the compound 1,3-
bis(N-carbazolyl) propane (1,3-BCP) (shown below).

H
H’
,
I
J/HZ. /H
I \H

Whereas dilute fluid solutions of NIPC exhibit only normal molecular


fluorescence, solutions of 1,3-BCP exhibit a new structureless emission band
which peaks approximately 5400 cm-1 to the red of the (0,O) band of NIPC
(Klopffer and Liptay, 1970; Klopffer, 1969b). The lower energy, structure-
less band has been attributed to intramolecular excimer emission. It is this
property which makes 1,3-BCP a useful model (see Fig. 12) for PVK since
solutions of the polymer also show a similar structureless emission band.
Values for the various rate parameters involved in intramolecular excimer
emission were arrived at by measuring relative intensities of monomer and
excimer emission in the absence and presence of dissolved oxygen. Johnson
(1974) has recently made both photostationary and transient measurements
on various carbazole “double molecules” in solution including 1,3-BCP;
1,3-bis[2-(N-ethylcarbazolyl)] propane (2,2‘-BCP) and l-[Z(N-ethylcarba-
zolyl)]-3-[3’-(N-ethylcarbazolyl)]propane (2,3’-BCP). Intramolecular excimer
emission is observed only for those compounds in which the carbazole groups
are symmetrically attached to the 1,2 positions of a propane chain. This
implies that the geometry of the excimer is essentially sandwichlike with the
two carbazole groups eclipsed. The thermodynamic parameters for forma-
tion of intramolecular eximers in l ,2-bis(N-carbazolyl) propane were also
determined to be A H = 2.76 kcal/mole, A S = 7.2 eu.
98 PENWELL, GANGULY, AND SMITH

28 26 24 2 2 . 1 0 ~cm.

Fig. 12. Fluorescence spectrum of solutions of NIPC (1) and BCP (2) in cyclohexane,
mole carbazole groupsjliter in the absence of air. The ordinate of curve 2 is extended by a
factor of 2.2, excitation. Reprinted with permission from the North Holland Publishing Co.
and the author, Klopffer (1969b).

Klopffer has studied both singlet and triplet energy transfer in films of
PVK (Klopffer, 1969a; Bauser and Klopffer, 1970; and Klopffer and Fischer,
1973). Singlet energy-transfer fluorescence in films of PVK has been shown
to be due to an exciton diffusion process. In amorphous aromatic polymers,
excitons are excited electronic states, randomly hopping from one side group
to the next; the side groups belong to the same or to different.polymeric
1
x103

3
E

Fig. 13. Fluorescence spectra of PVK in benzene solution (0.05 g/liter) saturated with ni-
trogen (l), air (2), and oxygen (3) together with the absorption of PVK in benzene (4) and
the solid-state fluorescencespectrum (5). The areas under the fluorescencecurves correspond
to the relative quantum efficiencies. Reprinted with permission from the American Institute
of Physics and the author. Klooffer (1969a).
POLY(N-VINYLCARBAZOLE) 99

chains. The range of the exciton is small compared to that observed in mo-
lecular crystals because of the presence of excimer-forming sites which act
as exciton traps. In PVK the concentration of excimer-forming sites is esti-
mated to be 10-3 mole/mole of carbazole residues. As a result, only excimer
rather than monomer fluorescence is observed in solid polymers like PVK.
Figure 13 shows the fluorescence spectra of PVK in benzene solution sat-
urated with Nz, air, and oxygen and the solid-state fluorescence spectrum.
Monomeric carbazole derivatives show phosphorescence in the wavelength
region between 410 and 500 nm (Klopffer and Fischer, 1973). PVK in dilute
frozen solutions also shows phosphorescence in this region (see Fig. 14).
In addition, delayed fluorescence from triplet-triplet annihilation is exhibited
(see Fig. 15). The latter depends on the molecular weight of the polymer;

Fig. 14. Phosphorescence spectrum (uncorrected) of 10-4 mole/liter NIPC in 2-methyl


tetrahydrofuran at 77 K. Excitation wavelength 295 nm. Reprinted with permission from
John Wiley and Sons, Inc. and the authors, Klopffer and Fischer (1973).

high molecular weight samples show very strong delayed fluorescence,


whereas in low molecular weight products phosphorescence prevails. Burk-
hart (1976) has suggested that the delayed fluorescence in PVK arises from
an annihilation process involving a mobile triplet and an immobile triplet
excimer.
The pbr-hnrescence of solid PVK is governed bv trans. CornParison
100 PENWELL, GANGULY, AND SMITH

ZI
L
C
0
I
c

EP
K
g

I 1
500 550 nm 6(
wavclcngth

Fig. 15. Phosphorescence and delayed fluorescence of PVK in 2-methyl tetrahydrofuran


(uncorrected), mole carbazole groups/liter ;emission slit 5 nm. Reprinted with permis-
sion from John Wiley and Sons, Inc. and the authors, Klopffer and Fischer (1973).

with low-temperature fluorescence and measurement of decay times suggest


that two traps are important for both singlet and triplet excitons. The
excimer-forming site is one of them. The maximum triplet excimer emission
is a t about 500 nm and the decay time is 7.9 t 1 sec (compared to 410 nm
and 43 nsec in the singlet system). The second trap emits in the same wave-
length region (decay time 14 sec) and corresponds to a second intrinsic singlet
trap emitting at 380 nm (decay time 12 nsec). The range of the triplet exciton
at 77°K is drastically reduced by those traps as in the case of the singlet
system.

VIII. PHOTOCONDUCTIVITY IN PVK : EXPERIMENTAL STATUS

Hoegl et al. (1958) first reported the photoconductive properties of poly-


vinylcarbazole (PVK) and pointed out its possible application in electro-
photography. They also indicated that the photosensitivity of PVK is en-
hanced by the addition of small amounts of dyestuff compounds and various
other organics such as carboxylics, sulphonics, quinones, aldehydes, and
ketones. These photosensitization effects and their role in electrophoto-
graphy will be discussed later on in this section. The fundamental under-
standing of the conduction mechanism in amorphous materials, including
polymers, is presently in its infancy and thus we will refrain from discussing
the processes of charge transport in PVK in any great length except to point
out to the readers a few scattered thoughts prevailing in the literature.
POLY(N-VINYLCARBAZOLE) 101

There are basically three distinct experimental modes for investigating the
transport of charge carriers generated by pulsed light (or any other means of
of generating free carriers in a photoconductor, viz., electron beam or
alpha particle) through highly resistive materials under an applied electric
field. First is the direct-contact measurement, developed by Spear (1969) and
used extensively by Markand Helfrich (1961) and Many and Rakavy (1962).
In this method a transparent electrode is deposited on the photoconductor
and direct contact is made to a constant voltage source [see Fig. 16(a)]. The
electric field polarity is determined by the charge-transport requirements,
i.e., electrons or holes. The external measurement in this mode is current
and thus this mode is often referred to as the current or J-mode.

TRANSPARENT

VOLTAGE CASE A
DETECTOR

i
TRANSPARENT
ELECTRODE

CONSTMlT
VOLTAGE CASE B
flDETECTOR1 1 SOURCE

CORONA
DISCHARGE TRANSPARENT
I. I I ELECTRODE

POTENTIAL CASE C
PROBE

Fig. 16. Three basic experimental configurations, cases A-C, for the pulsed light photo-
discharge measurement. Reprinted with permissionfrom the authors, Seki and Batra (1971).
Copyright by International Business Machines Corporation.

The second configuration is a modification of the above case where the


photoconductor is essentially in series with one or two insulating dielectrics
and a transparent electrode is deposited at the top of the system [see Fig.
16(b)]. Contact with the constant voltage source is made prior to injection
of the carriers and the external measurement is current. This technique
has been developed by Many (1965) and Tabak (1967).
Finally, in the floating surface case [see Fig. 16(c)] the photoconductor
is charged by a device such as a corona or other charged-beam source and
avoids any direct contact with a voltage source. The voltage discharge which
102 PENWELL, GANGULY, AND SMITH

is due to photoconduction is measured by a transparent surface potential


probe which is coupled to the sample with a negligible capacitance. The ex-
ternal measurement in this case is the voltage drop across the photoconductor
and is often referred to as voltage or V-mode.
The first systematic study of the photoconductive behavior of thin films
of PVK in the V-mode, both for PVK-generated and Se-injected carriers,
was made by Regensburger(1 968). We will discuss the photoconducting behav-
ior of PVK in the PVK-generated case and the injected-carrier cases sep-
arately.

A. Photoconductivity in PVK
Lardon et al. (1967) have shown that the xerographic gain (XG),defined
as the ratio of the initial discharge rate (dV/dt),=,to the rate at which the
discharge will take place if one charge carrier per incident photon were ex-
cited and moved across the photoconductor, eFd/c (A) where e = charge,
F = light flux, d = thickness of the PIC, and E (A) = wavelength-dependent
dielectric constant, follows the absorption spectrum very closely (Fig. 17).
This phenomenon is physically conceivable because the wavelength-depen-
dence of the dielectric constant enters both XG and absorption spectra in an
analogous way. The fluorescence spectrum for an excitation wavelength of

Roton ennpy (cV)


45 4 0 3 5 30

Poly (N-vmyl corbozole)

I"
50

0
-CHrHC-
40 .
' 0
0

"
z,P
30
?
X

2 0 6

Flwrescence 10
sDtctrum
,_..._
,'.....-.. ........-.., ....
'\

,.y, '
.C

*,

2500 Moo 300 4ooo 4500

Fig. 17. Absorption spectrum, luminescence spectrum, and the dependence of photo-
conductivity for a 7.6 p PVK layer at 630-V potential. The illuminated surface is positive.
Reprinted with permission from Pergamon Press, Ltd., and the author, Regensburger
(1968).
POLY(N-VINYLCARBAZOLE) 103

of 2537 A is also shown in Figure 17. In the electrical measurements of Re-


gensburger (1968) the following conclusions were drawn :
(1) The initial rate of discharge (dV/dt)(,=,as a function of the initial
applied field EO for hole conduction (shown in Fig. 18) seems to obey the
empirical relation

Fig. 18. Electric field dependence of hole phofocurrent and gain of PVK layers directly
excited by 2.2 x 1013 photons/cm2 sec at 3450 A. Reprinted with permission from Perga-
mon Press, Ltd., and the author, Regensburger (1968).

'rI)I
d dt t=O
= 1.6 x 10-3 (p/V sec) E 2

for a photon flux of 2.2 x 1013 photons/cm2 sec.


(2) The xerographic response for electron transport for 253 A excitation
is about 200 times smaller than that of hole transport response.
(3) The initial discharge rate for hole transport for 3450 h; excitation is
linearly dependent on light intensity up to 5 x 1013 photons/cm2 sec.
(4) The response time T,, assumed to be the same as the transit time of the
fastest carrier, is inversely proportional to the cube of the applied field (Fig.
19):
4
10

10 1 , . . , , . I , . , . , , , . , , . , . ,

Fig. 19. Reciprocal response time as a function of applied field. Reprinted with permis-
sion from Pergamon Press, Ltd., and the author, Regensburger (1968).

d
- = 5.0 x
T,

(5) The initial pulse current (in arbitrary units) is proportional to the fifth
power of the applied field (Fig. 20).
(6) Finally, there is an exponential increase in the hole mobility with tem-
perature at a constant field. This seems to indicate a field dependence of the
activation energy varying between 0.4 and 0.7 eV.
Points (1)-(6) are the main “observed” results of the experiment. How-
ever, Regensburger has also drawn other conclusions and has estimated
several relevant entities. For example, recalling the definition of the drift
mobility u
, as velocity per unit field, one can derive an electric field dependence
--f mobility from Figure 19:

d/Tr
,u = -= 5.0 x
E

The above square-law electric field dependence of the mobility and the con-
stant factor associated with this law will be discussed elsewhere in this
section. In a similar attempt, the initial pulse current can be related to the
quantum efficiency of charge generation v :
POLY(N-VINYLCARBAZOLE) 105

+ 7.6pWK
a 3 0 p PVK

I
1
I0 10 Io2 lo3
E, (VIP)

Fig. 20. Initial pulse current as a function of applied field in 7.6-,u and 30-,u PVK layers
(two different light intensities were used). Reprinted with permission from Pergamon Press,
Ltd., and the author Regensburger (1968).

JO = qFe,uE/d

Since JO cc E5 and p cc E2, one can easily deduce the relation q oc E2. This
is consistent with the observations that (1)
e
'(C)I
d dt t=O = qF; cc E2

and (2), the initial pulse height AV is proportional to E 2 :


e d
A V = qF-- - cc E2,
E A
106 PENWELL, GANGULY, AND SMITH

where A is the sample area illuminated.


A recent careful reexamination of PVK-generated carrier transport by
Pai (1970) shows that the transient voltage (pulse height) has essentially an
exponential square-root electric field dependence. Although the effective
transit time is strongly dependent on the applied electric field and the thick-
ness of the film, the shape of the transient signal is exponential in time within
1 to 100 V/p range. The functional form of the field dependence of the
transit time was found to be Tr cc exp(-yE1/2). Measurements of Tr for
various thicknesses confirmed a linear dependence of Tr on thickness as
shown in Figure 21. It was pointed out in this paper that if one confines the

100 000 I, I I I I I

I
1: -' 2
1
3
I
4
I
5
I
6
I
7 8

IO-~ CcG)
Fig. 21. Inverse effective transit time (Te@ vs. square root of the applied field for four
film thicknesses. Reprinted with permission from the American Institute of Physics and
the author, Pai (1970).

measurement in a limited region of applied field, one may be prompted to


draw fortuitous conclusions regarding the forms of field dependence of tran-
sit time, viz., linear field-dependence (Szymanski and Labes, 1969). The field
dependence of the number of charge carriers that have transitted through the
sample for different thicknesses is depicted in Figure 22. As can be seen from
this figure, the number of carriers transitted through the sample in the high-
POLY(N-VINYLCARBAZOLE) 107

. 3x lo-‘
A - 1.44/&

%---
0-

3XIO-~

F
*

3
W
z
w
uU
U
W

Fig. 22. The product of the capacitance of the film and the amplitude of the transient
signal vs. square root of the applied electric field for four film thicknesses. Reprinted with
permission from the American Institute of Physics and the author, Pai (1970).

field region is independent of the thickness of the film and increases monoto-
nically with field. These two distinct behaviors are attributed to the absence
of any bulk trapping in the high-field case and the field dependence of the
quantum generation, respectively. The effect of bulk trapping in the low-field
case is demonstrated in Figure 23, where temperature variation is also in-
cluded. The change in slope in the logarithmic-square-root plots in these
curves separates the high-field, no-trapping region from the low-field, trap-
ping region. Finally, the thermal-activation energy calculated from an inverse
transit time versus temperature plot is found to be 0.36 eV. This value of the
activation energy can be calculated from shifting of the slope breaking point
as a function of temperature and field.
Pai (1970) has also studied the thermally stimulated current from the traps
in PVK. Figure 24 shows current versus temperature data for four different
voltages. The salient features of this experiment were the reproducibility of the
low-temperature peak and the field-dependent position (in temperature) of the
current peak. These observations led to the conclusion that the trap release is
influenced by the applied electric field as well as by the temperature.
108 PEWELL, GANGULY, AND SMITH

- - HEATING RATE 0.043'/*c


-0+
I I
FILM THICKNESS 5 MICRON
--- 4ov [
t 6.0-
- ---
W
a
a -
20v
IOV I-
POLY(N-VINYLCARBAZOLE) 109

B. Photoconductivity in Selenium-Sensitized PVK

The transport of the injected charge carriers from selenium into PVK was
studied systematically by Regensburger (1968). In this study a thin (-0.3 p)
sensitizing layer of amorphous selenium was evaporated directly onto thick
(- 10 p) PVK layers which were already solvent-coated onto polished al-
uminum. The XG in this configuration is quite high for the extended wave-
length range (visible to uv) and is comparable to that of selenium itself (see
Fig. 25).

I0

-1
10

-2
10

B
0)

u
c
0

g
e -3
; I0 -0 8 . 0 Selenium
~
o---* 0.3pSt/7.6~PVK
+-.-+ 7 . 6 PVK
~

-4
I0

*
3400 3800 4200 4600 5000 5400 5800 6

Wovelenqth tdJ

Fig. 25. The dependence of gain on wavelength for PVK, selenium and a thin overcoating
of selenium on PVK. Reprinted with permission from Pergamon Press, Ltd., and the author,
Regensburger (1968).
110 PENWELL, GANGULY, AND SMITH

10-

.f-*
../.
01 -
E
D

t ! .
r
a
: .
?
X

001 -
+----+ 0 3pSe/ 7 6 p P V K
- 3 03pSe/12 5 p P V K

1 I
Oool,
I0 I00

Fig. 26. Dependence of photocurrent and gain on field for a series of Se/PVK composites
with progressively thicker PVK. Reprinted with permission from Pergamon Press, Ltd., and
the author, Regensburger (1968).

Figure 26 shows the XG against the applied electric field for the composite
structure for different PVK thicknesses. The sharp threshold field (about
6 V / p ) was interpreted by Regensburger as a barrier to injection of carriers
from sensitizer Se to PVK. Subsequent careful studies (Mort, 1972; Mort and
Neilson, 1972) have established that there is no indication of any significant
barrier to injection and that the field threshold can be interpreted as being due
to trap-free space-charge-perturbed currents in PVK.
Figure 27 gives Mort’s data (1972) of current versus electric field for differ-
ent light intensities. For the highest light intensities, the transient current is
proportional to -IT4 and is essentially light-intensity independent. With the
decrease in light intensity there is a small but definite decrease in the slope
of the curve, particularly for lower fields.
Figure 28 gives the variation of the reciprocal transit time against field.
The E2.5field dependence found by Mort (1972) is somewhat weaker than
that of Regensburger (1968). From Figures 27 and 28, Mort concluded that
111

Fig. 27. Current density vs. electric field. The electric field is determined from the total
applied voltage divided by the PVK thickness. Reprinted with permission from the Ameri-
can Institute of Physics and the author, Mort (1972).

the magnitude of the transient current observed in the low-field region was
very close to the maximum bulk current that a corresponding thickness of
PVK could sustain because of space-charge limitations. The physical mech-
anism leading to the space-charge-perturbed current in PVK is the accumula-
tion of one -CV of charge at the interface for times much longer than the
transit time of holes in selenium. This essentially reduces the field to zero at
the generation surface until the original surface charge is restored only after
the circuit time constant. Because of the large dispersion in the mobility of
holes and the strong field dependence of mobility in PVK, the trailing edge of
112 PENWELL, GANGULY, AND SMITH

I o4 1 I

lo3

2
r10
0
0
v)
I

-z
IC

I .c

0.1
10' 10' lo5 106
E(v/cm)
Fig. 28. Reciprocal of fastest transit time vs. electric field. Reprinted with permission from
the American Institute of Physics and the author, Mort (1972).

the charge injected into PVK from selenium moves very slowly. Thus the
charge generation and transport in selenium will be greatly reduced for times
longer than the transit time in selenium and shorter than that in PVK.
The above mechanism of space-charge-perturbed current in PVK is con-
firmed by experiments based on accelerated transit time techniques (Mort,
1972). In this low-field experiment a second pulse of light is generated
before the CV of charge generated with the previous pulse has moved away
from the Se-PVK interface. Since the trailing edge (which has a substantial
part of the injected carriers) has not moved appreciably away from the inter-
face at low fields, the field at the generator surface is almost zero. Thus,
essentially no selenium response is seen (because of the field dependence of
POLY(N-VINYLCARBAZOLE) 113

generation in selenium). The signal can be restored either by waiting for sever-
al transit times or by increasing the field across the PVK, encouraging the
faster passage of the charge through PVK.
A few simple but informative experiments carried out by Mort (1972)
deserve attention. In order to establish convincingly the presence of trap-
free space-charge-perturbed current in PVK, the transient and steady-state
experiments are conjugated. The basic idea contained in this experiment is
that if the space-charge-perturbed current in PVK is an indirect manifestation
of low mobility of carriers in PVK, then simultaneous steady-state and
transient currents must add up to the space-charge-perturbed current.
Figure 29 verifies this criterion of trap-free space-charge-perturbed current
in that the transient current in the case of simultaneous illumination is re-
duced almost by an amount equal to the steady-state current. It is worth
pointing out that the above discussions are based on the assumption that the
PVK is free of any appreciable deep-trapping centers in the bulk or any sub-
stantial carrier loss at the interface because of recombination and/or a variety
of other reasons.

T I M E (IOrnsec/div.)
( 0 )

?
TIME ( lOrnsec/div.)
(b)
Fig. 29. (a) Trace of TFSCPC transit pulse in PVK. (b) Trace of TFSCPC produced by
simultaneous dc and excitation. The dc current was adjusted to be one division. Reprinted
with permission from the American Institute of Physics and the author, Mort (1972).

Three independent methods were employed by Mort (1 972) to determine


the absence of trapping in the bulk Se:PVK system and/or loss of carriers at
114 PENWELL, GANGULY, AND SMITH

the interface. First, the area under the current curves for Se-transit and PVK-
transit were measured graphically for different applied fields. No substantial
difference was found to account for the injection threshold field of 6.5 V/p.
The second crucial test employed to overrule any bulk trapping was the meas-
urement of the temperature dependence of the light-intensity-independent
transient currents. The results are shown in Figure 30 for -30 to 60"C tem-+

2.9 3.1 3.3 3.5 3.7 3.9 4.1


1 0 3 (OK)-'
1
'

Fig. 30. Reciprocal of fastest transit time and current density vs. reciprocal temperature.
Reprinted with permission from the American Institute of Physics and the author, Mort
(1972).

perature range. The thermal activation energy for hole transport in PVK,
determined from the slope, is about 0.5 eV and is independent of the applied
field. The nearly identical slope of the two curves in Figure 30, in conjunction
with the reproducibility of data with respect to heating and cooling, rules
out the possibility that any barrier effects or release from any interfacial traps
are responsible for the loss of injection efficiency at a threshold field. It is to
be noted that these temperature-dependence experiments were not feasible
at fields near or below the threshold field of 6.5 V/p and thus, strictly speak-
ing, the conclusions are interpolated. Third, the method of interrupted tran-
sits (Gibbons and Spear, 1966) is adopted to find out the existence as well as
lifetime of deep traps in the bulk of PVK, if any. The basic idea in this experi-
ment is to interrupt the transit of the carrier through PVK by shorting the
sample so that the field across the sample is zero. This effectively makes the
transit time infinitely long compared to that of the deep-trapping lifetime and
POLY(N-VINYLCARBAZOLE) 115

thus allows trapping to occur. When the field is restored after a certain time,
the remaining free carriers move with a drift velocity. The reduction in ob-
served current as a function of interruption time yields information concern-
ing the number of deep-trapping centers and deep-trapping lifetimes.
The currents observed (Mort, 1972) in this experiment in the Se:PVK
system are shown in Figure 31. The lower value of current following the inter-
ruption indicates some loss of carriers. Another noticeable effect in the inter-
rupted case is the reduction in drift mobility, which is not well understood.
It should be kept in mind that these measurements involve very small cur-
rents (- 10-11 A), which are comparable to the noise level of the circuit
with the RC constant limited by possible maximum series resistance. The
procedure of using differential current-integrated values for interrupted and
usual transits may involve considerable error because of very long (order of
hours) transit times of trailing carriers in low fields in PVK. Moreover, there
may be some loss of carriers from redistribution of charge during the inter-
rupted time period.

- t 1

I . . 1 I I I I I J
TIME (5rec/diu.)

FIELD 6.71 103V/cm

Fig. 31. Results of interrupted transit experiment on PVK. (a) Current pulse without
interruption and (b) current pulse after transit was interrupted for 4.5 sec. Trace (b) has
been drawn leaving out the interruption period. Reprinted with permission from the
American Institute of Physics and the author, Mort (1972).

The 6.5 V/p threshold field in the observed XG vs. field curve (Fig. 26) is
explained in terms of that constructed on the basis of the maximum current a
10-p-thick PVK can carry and the amorphous selenium generation-limited
current. In Figure 32 the trap-free space-charge-limited current (TFSCLC)
is shown for low fields for a 10-p PVK:Se structure. The XG data were also
determined experimentally from initial rate of change of voltage drop meas-
urements in a xerographic mode (V-mode). A XG of 0.8 was found at a 10
V/p field with 2 x 10l2 photons/cm2 sec flux. This data point is employed to
determine the Se generation-limited current because at this high field the cur-
116 PENWELL, GANGULY, AND SMITH

lo3 lo4 lo5 106


E(V/cml

Fig. 32. Comparison of measured xerographic gain (XG) vs. field curve with that con-
structed using a composite of TFSCPC curve, measured on the same sample, and the
amorphous Se generation-limited current curve. The latter was computed for the light
level used in the xerographic measurements with the experimentally observed gain of 0.8
at 106 V/cm. Reprinted with permission from the American Institute of Physics and the
author, Mort (1972).

rent should be generation-limited in a Se-PVK system. Assuming that the


generation efficiency in selenium is proportional to Eom5, this high-field-
current value determines the proportionality factor and thus allows one to
calculate the Se generation-limited currents for lower field values. The Se
generation-limited current thus obtained corresponds approximately to a
calculated steady-state trap-free current, assuming quantum efficiency of 30 %
for a photon energy of 3 eV a field of 10 VIP. It is noteworthy that in Figure
32 diverse experimental data points are compiled, viz., TFSCLC data are
current-mode data whereas the XG data are voltage-mode data ; TFSCLC
data are the same for the time of measurement within the transit time of the
fastest carrier or in the steady state, whereas the Se generation-limited current
POLY(N-VINYLCARBAZOLE) 117

data will be quite different in these two cases. Therefore, to make the com-
parison meaningful, the XG data must be taken within the transit time of the
fastest carrier.
From the foregoing analysis we see that there canbe two distinct kinds of
bulk limitations in a low-mobility material. First, the usual “range limitation
of Kind I” arises primarily from deep traps in the material and is stated by the
condition pzE < L, where L is the sample thickness. z is the mean free time
of the carrier before it is permanently trapped. The second is “range limita-
tion of Kind 11” in which, because of low mobility as well as dispersion in
mobility, a large fraction of the carriers have not sufficient time to transit
across the material before the initial xerographic measurements are made.
Mort and Neilson (1972) have attempted to observe directly the accumula-
tion of charge at the selenium-PVK interface using a double relay technique.

10w 10
10‘ 10’ 106
EWcm)
Fig. 33. dV/dt vs. field for a 22-p PVK film overcoated with 2 p of amorphous Se at five
different light intensities. The +’s represent initial dV/dt values measured at the highest
flux of 3.1 x 1014 photons/cm2 sec. The other four curves were obtained by differentiation
of single discharge curves. Reprinted with permission from the American Institute of Phy-
sics and the authors, Mort et al. (1972).
118 PENWELL, GANGULY, AND SMITH

Fig. 34. Plot of dV/dt vs. time for the Se:PVK sample measured at the four light fluxes.
Reprinted with permission from the American Institute of Physics and the authors, Mort
et al. (1972).

The basic principle involved in this measurement is to vary the time CQ (required
for initial surface charge Q to accumulate at the interface) and the transit
time of the fastest carrier in PVK, tT. This is done by varying the dark con-
ductance of the selenium by changing the light intensity and by varying the
bias applied across the Se:PVK system. If t~ << tT for a certain high light
intensity and low bias, a time-resolved transit in PVK will be observable. For
tQ I ZT the transient current will trace out the buildup of charge at the inter-
face and the transit of carriers through PVK will not be resolved. For very
high field when transit through PVK is almost selenium-emission-limited, tQ
becomes very large and thus no current decay can be observed because there
is no charge accumulation at the interface. The above conjectures have been
verified by actual experiment (Mort and Neilson, 1972; Regensburger, 1968).
Except for the early work of Regensburger, the studies on Se:PVK systems
described thus far are in the current or J-mode. In what follows we shall de-
scribe the effect of space-charge perturbation, arising from the range limita-
tions of Kind I and/or Kind I1 on xerographic behavior (V-mode) of a Se:
PVK system (Mort et al., 1972). The dependence of the discharge rate (dV/dt)
on the applied field and time are shown in Figures 33 and 34, respectively, for
different light fluxes. At low fields a common value of dV/dt is found with an
POLY(N-VINYLCARBAZOLE) 119

E4 variation for all light intensities. This E4 dependence is, as before, the man-
ifestation of trap-free space-charge-limiteddischarge of PVK. In conjunction
with the modified Child's law, V4 dependence of discharge rate, which is
proportional to current in the system, leads to an E4 dependence of mobility.
Using the modified Child's law,

- -
one gets from Figure 33 p 4 x lo-' cm2/Vsec for E 3.3 x lO5V, L 20
p, and K - 3.47. An independent verification of the order of magnitude of
-
the mobility can be obtained from Figure 34 in the following way. The inter-
section of the extrapolation of the constant (dV/dt) line and the high-flux,
long-time asymptotic (dV/dt)line yields the value of transit time as 2 x 10-2
sec. This corresponds to a mobility of 3.6 x lo-' cm2/V sec at the initial field
of 3.3 x 105 V/cm. This is in good agreement with the values of mobility
obtained from Figure 26 and also from the time-of-flight measurements (Mort,
1972). Figure 35 shows the discharge characteristics of Se:PVK under finite
exposure of 8 msec for two different light fluxes of 3.1 x 1014 and 3.1 x 1013
photons/cm2 sec. At the higher flux the transport of carriers in PVK is space-
charge-limited and thus discharge is not completed during the exposure time,
whereas for low flux the surface potential is almost constant after the exposure
time.

I I
-
-
-
-
-
2
w
c
0 300- -
a
E200- -
2
a
3 100- -
v)

0 I I I I

Fig. 35. Discharge curves of Se:PVK sample at fluxes of 3.1 x 1014 and 3.1 x 1013
photons/cma sec, respectively,following an exposure of 1/60 sec. Reprinted with permission
from the American Institute of Physics and the authors, Mort et al. (1972).

The contrast potential obtained by subtracting the two curves of Figure 35


is shown in Figure 36. The maximum contrast potential obtained after about
half a second is -300 V for an input density of one with a flux of 3.1 x 104
120 PENWELL, GANGULY, AND SMITH

I 1 1 I

0
0 0.I 0.2 0.3 0.4
J0.5

TIME(sec)

Fig. 36. Growth of contrast potential vs. time for an optical density of 1.0 with a white
background flux of 3.1 x 1014 photons/cmz sec. Reprinted with permission from the
American Institute of Physics and the authors, Mort et at. (1972).

photons/cm2 sec. The sample was initially charged to -800 V. The effect of
space-charge perturbation on the time development of contrast potential for
different light intensities and finite exposures but with equal integrated flux
input to the system is shown in Figure 37. The zero-input density fluxes in
Figure 37(a)-37(c) are 3.1 x 1014, 3.1 x 1013, and 3.1 x 1012, respectively.
For lowest intensity and highest exposure time (to have equal integrated
flux), the contrast potential measured immediately after the exposure and
maximum contrast potential attained during the discharge almost coincide,
whereas for highest intensity and lowest exposure time the two measurements
of contrast potentials differ considerably. The latter is a direct consequence
of space-charge limitations because the transit of the injected charge is not
completed during the exposure (8 msec). The transit time of the fastest car-
riers in these fields is about 2 msec.
Carnes and Warter (1 972) have made measurements of the electric current in
PVK associated with the return to thermal equilibrium after photoexcitation.
In this measurement a thin layer of PVK was put between blocking electrodes
and was photoexcited by weakly absorbing light. After the light was removed,
voltages were applied and the time evolution of the resulting current in excess
of the capacitative charging current was measured. The variations of this dif-
ferential current on electric field, thickness, exposure, and temperature of the
sample were monitored. The conclusions drawn from these observations were :
(1) the hole range in PVK is greater than 4 x 10-6 cm2/V; (2) the charge
carriers observed are stored only at the surface and either recombine there
or transit completely across PVK when released.
POLY(N-VINYLCARBAZOLE) 121

2
Y ---I
200
/ \
-\

EXPOSURE (PHOTONS)

Fig. 37. Contrast potential vs. exposure curves for an optical density of unity at different
white background fluxes. The initial potential in each case was 820 V. White background
flux: (a) 3.1 X 1014 photons/cmZ sec, (b) 3.1 x 1013photons/cmz sec, (c) 3.1 X 10l2pho-
tons/cm2 sec. Reprinted with permission from the American Institute of Physics and the
authors, Mort et al. (1972).
122 PENWELL, GANGULY, AND SMITH

C. Photoconductivity in Dye-Sensitized PVK

As pointed out at the beginning of this section, the photoresponse of poly-


cyclic aromatics and heterocyclic polymers can be appreciably enchanced by
the addition of a variety of electron acceptors like acids, acid hydrides, esters,
ketones, aldehydes, quinones, nitro compounds, cyano compounds, halogen
compounds, azo compounds, and many dye sensitizers: Nile blue A dye salt,
rhodamine B, rose Bengal, malachite green, etc. (Hoegl et al., 1958, 1972;
Lardon et al., 1967; Ikeda et a]., 1969, 1970; Peredereeva et al., 1970; Schaf-
fert, 1971 ;Melz, 1972; Queener, 1971 ;Hughes, 1972; Ohno et al., 1972; Tada
et al., 1972; Okamoto et al., 1968, 1973; Meier et al., 1972, 1972a, b ; Weigl,
1972; Sato, 1972; Gill, 1972; Klopffer and Willicks, 1968; Smith and Weigl,
1965; Meier, 1963, 1965, 1969; Bourdon and Schnuriger, 1967; Oka et al.,
1964a, b; Hayashi et al., 1966; Bauser and Klopffer, 1970a; Epping, 1970).
These photosensitizers not only extend the sensitivity of PVK to the visible
region but also increase the sensitivity of PVK in its near-uv intrinsic absorp-
tion region. As an example, Figure 38 shows the effect of 1 mol % doping with
2,4,7-trinitrofluorenone-(9) (TNF), l-chloro-2,4,5-trinitrobenzene(picryl

0.000 a.OO0 25.000 30.000

Y. cm-'

Fig. 38. Absorption spectrum of PVK in benzene solution and of PVK films with various
dopants in equal molecular ratio. Reprinted with permission from Pergamon Press, Ltd.,
and the authors, Hoegl et al. (1972).
POLY(N-VINYLCARBAZOLE) 123

A . rm
850750 650 550 450 350

TNFh w k t i

10 md -XTNF //
w‘
w
-8
-
0

/ I md -XT NF p. I

+
Fig. 39. Absorption spectra of PVK TNF films and of TNF in ethanol solution.
Reprinted with permissionfrom Pergamon Press, Ltd., and the authors, Hoegl et al. (1972).

chloride) (PC), tetrachlorophthalic anhydride (TCPA), and 1,Zbenzoanthra-


quinone (BA) on the absorption spectrum of PVK film and PVK in benzene
solution. Figure 39 shows the variation of the absorption spectrum of TNF-
doped PVK with different concentrations of TNF. The appearance of new
absorption bands (cf. Fig. 39) at longer wavelengths which are absent in the
other components prompts the hypothesis of intermolecular charge-transfer
absorption (Lardon et al., 1967; Lell and Weigl, 1964; Sharp, 1967). The
XG as a function of wavelength for hole transport for various dopants in PVK
is shown in Figure 40, along with that in PVK for hole as well as electron
transport. There is a close similarity between the absorption spectrum and XG
spectrum as pointed out in Figure 17. The dopants generally do not appreci-
ably increase the XG of PVK in its intrinsic absorption region (“chemical”
sensitization). The increase in sensitivity by doping is mostly due to an exten-
sion of photoresponse and a strong increase in absorption in the visible region
(“optical” sensitization).
The photoinduced discharge characteristics (PIDC) for continuous illu-
mination of 2,4,5,7-tetranitro-fluorenone-(9)(TENF)-doped PVK is shown in
Figure 41. The discharge rate increases monotonically with increasing TENF
concentration.
The number of sensitizers which improve the photoresponse and photo-
conductivity of PVK is quite large, and we refer the reader to the work of
Oramoto et al. (1973a) and Wagner and Gasner (1970) for greater detail. The
basic mechanisms of dye sensitization of PVK and other organic photocon-
ductors are still under discussion, and we shall touch upon them very briefly.
124 PENWELL, GANGULY, AND SMITH

DTF
IC

\ P V K + O - I ~ O I - Y .BBOT

.-c I 0
8
-% NB

10-

I0 350 450 550 650 750

A. nfn

+
Fig. 40. PVK various dopants; gain as a function of wavelength. Reprinted with
permission from Pergarnon Press, Ltd., and the authors, Hoegl et al. (1972).

Time

Fig. 41. Discharge characteristicsof PVK films with increasing TENF dopant concentra-
tion. Reprinted with permission from Pergamon Press, Ltd., and the authors, Hoegl et al.
(1972).
WLY(N-VINYLCARBAZOLE) 125

As pointed out at the beginning of this subsection, the occurrence of new


bands in electron-acceptor-doped PVK is often interpreted as being due to a
charge transfer between donor and acceptor (Bourdon and Schnuriger, 1967)
and that sensitization is a result of this mechanism (Sharp, 1966, 1967a; Aka-
mot0 and Kuroda, 1963; Kuroda et al., 1964). Two distinct mechanisms
of sensitization have been, in general, accepted in the literature, viz., an ener-
gy-transfer process and a charge-transfer process. In the energy-transfer pro-
cess the excitation energy is transferred from the light-absorbing molecule of
dye to the carriers trapped in the local levels of PVK. Subsequent excitation
of these trapped carriers into conducting bands results in photosensitivity.
In the charge-transfer mechanism the process of sensitization consists of crea-
tion of an electron-hole pair in the dye and transfer of the electron or hole
from the dye to the conducting states of the parent photoconductor. Since
some of the dyes studied do not possess photoelectric sensitivity, the first
mechanism accounts for photosensitization processes.
The observed change in the intrinsic photoelectric sensitivity of a polymer

I
I
I
? I

i
0
X
U
L
+
C I
?L!
3
V
0
c
0
c
a

400 500 600

Wavelength, nm
+
Fig. 42. Action spectrum of photoconductivity of sensitized PVK. (1) PVK methylene
+ +
blue. (2) PVK methylene blue TCNE. Reprinted with permission from Pergamon
Press, Ltd., and the authors, Meier et al. (1972).
126 PENWELL, GANGULY, AND SMITH

in the presence of a dye is not yet completely understood. It is possibly con-


nected with the quantum efficiency of charge and/or exciton generation in the
dye, the quantum efficiency of tunneling of charge and/or exciton from the
dye into the available states of polymer and the quantum efficiency of juxtapo-
sition of carriers in the polymer conducting states leading to free carrier
generation for effective photoconduction. There are attempts to separate the
magnitudes of different quantum efficiencies for apparently eiectron-trans-
ferring dyes (carbonium salts, benzopyrylium salts, and crystal violet) in
PVK. In general, the effective quantum efficiency of charge generation and its
electric-field dependency in a sensitized photoconductor is governed primarily
by that of the parent photoconductor.
Meier et al. (1972a,b) have observed an additional new peak in the action
spectrum of a three-component system consisting of PVK/acceptor/dye. This
new band is quite distinct from the polymer, the dye, and any combination of
charge-transfer absorption bands.
+
Figure 42 shows the action spectrum of (PVK methylene blue) and (PVK
+
+ methylene blue TCNE). Figure 43 shows an analogous spectrum of
+ +
PVK pinacyanol TCNE and absorption spectrum of pinacyanol. The
new band at 490 nm in Figures 42 and 43 lies between the polymer and dye
band and is independent of the nature of the dye because there is no observable

\ I
\ I
\ I
\II
POLY(N-VINYLCARBAZOLE) 127

change in the position of the new band going from methylene blue to pinacya-
nol, crystal violet, or many other dyes. On the other hand, the position of the
new peak depends on the nature of the added electron acceptor. For example,
the peak shifts from 490 to 410 nm by changing the electron acceptor from
TCNE to TCNQ. However, the intensity of the additional band varies with
concentration of the dye; the peak intensity of the new band increases with
an increase in dye concentration whereas the dye-sensitization band intensity
decreases. Some of the obvious possibilities of the origin of this new band can
be ruled out easily. This band at 490 nm cannot be related to the charge-
transfer (CT) band of PVK-TCNE because this CT band is situated at 600
nm. Moreover, the position of this 490 nm band cannot be related either to
the dye-acceptor CT band because the position of the band does not depend
on the nature of the dye absorption bands or to the absorption band of the
neutral acceptor for TCNE because the neutral acceptor band is at 280 nm.
The hypothesis put forward to account for the 490 nm band is the formation
of acceptor radicals of TCNQ, TCNE and subsequent indirect sensitization
of PVK by the dyes on the basis of electron-transfer mechanism through ac-
ceptor radicals. The possibility of appropriate Fermi resonance between the
energy levels of the polymer, the acceptor, the dye, and their charge-transfer
states cannot be ruled out as a possible explanation of the occurrence of the
490 nm band.

D. Photoconductivityof Charge-Transfer Complexes of PVK


In this subsection we will discuss in some detail the photoconductivity of
highly loaded PVK with electron-acceptor organic materials such as T N F
(Shattuck and Vahtra, 1969; J-P. Montillier, 1974). Any electron-acceptor
organic molecule should, in principle, form a charge-transfer complex at
high loadings. Chlorine-, hydroxy-, and nitrosubstituted anthraquinones in
molar ratio 0.1-0.5 in PVK showed good photoresponse, but none were as
good as the 1 :1 ratio PVK:TNF charge-transfer system (TPC). The photocon-
ducting behavior of this system has been studied in IBM and Xerox labora-
tories (Lardon et al., 1967; Schaffert, 1971; Hughes, 1972; Gill, 1972). The
photosensitivity, measured as the reciprocal of the exposure time required to
reduce the surface potential to one-half its initial value, of 1 :1 molar ratio
PVK:TNF for positive and negative charging is shown in Figure 44. Concen-
trations of T N F beyond 1 :I molar ratio do not produce any improvement in
sensitivity. The electrical and optical properties of 1 :1 molar ratio PVK:TNF
systems have been studied extensively and we will only compile a few relevant
results for this system.
The dark- and photocurrents as a function of applied voltage and light
intensities are given by Figures 45 and 46. For higher voltages the photocur-
rent follows an approximate P . 3 power law as can be seen from Figure 45.
The photoconductive gain, defined loosely as the ratio of photocurrent to
light intensity as a function of applied voltage, is a rather steep function of
voltage across the photoconductor-an increase of 200-fold, from 50 to 900 V
128 PENWELL, GANGULY, AND SMITH

I Concentration in molos TNFI moles cerbatole


Fig. 44. Photosensitivity as a function of TNF concentration in PVK. Reprinted with
permission from the author, Schaffert (1971). Copyright 1971 by International Business
Machines Corporation.

across a 20-,u sample. The photocurrent increases as the 0.75 power of light
intensity and is about linear for all light intensities for low voltages and lin-
ear beyond 25-30 ,uW/cm2 for higher voltages.
The dark resistivity as a function of different conducting substrates and
applied field is shown in Figure 47. The samples coated on anodized alumi-
num have a higher dark resistivity in the high-field region, providing a slower
dark decay of the surface potential and higher charge acceptance. The vari-
ation of the dielectric constant at a frequency of 200 Hz as a function of T N F
concentration is shown in Figure 48. The dielectric constant has a maximum
of about 3.2 at 0.5 molar concentration of T N F and is about 2.6 5 0.1 for a
1 :I molar ratio. The dielectric constant of pure PVK at 200 Hz is 2.8 k 0.06
(cf. Fig. 48), which is quite distinct from the value of 3.75 0.4 at 10 Hz
reported by Regensburger (1968).
The enhancement of the amount of absorption in the visible part is pri-
POLY(N-VINYLCARBAZOLE) 129

o Dark
0 l.OpWIcrn?
A 18.3pWIcm2
x 50.0pWlcm2
V 150.0pWlcm2

1 : 1 TPC on anodixd
Al. thickncc\= 201" rn

0 200 400 600 800 1000 1200

Applied potrntial i n volts

Fig. 45. Current-voltage curves for a typical 1:l TPC sample. Data were obtained by
screen-controlled corona charging. Reprinted with permission from the author, Schaffert
(1971). Copyright 1971 by International Business Machines Corporation.

marily due to the charge-transfer bands of the PVK:TNF complex. The ab-
sorption cutoff at about 425 nm (cf. Fig. 49) agrees well with T N F single-
crystal data (-450 nm).
The variation of charge acceptance, i.e., the maximum amount of charge
that can be applied to and retained on a photoreceptor in the dark before
dielectric breakdown takes place, with different concentrations of TNF, is
shown in Figure 50.
The distinct differences in the behavior of the PVK:TNF system upon
negative and positive charging corona are clear in Figure. 50. The superior ne-
gative-charge acceptance of a 1:l molar ratio PVK:TNF system has been
exploited in making PVK :TNF photoreceptors for the IBM copier machine.
The variations of dark decay rates, defined as the rate at which a charged
photoreceptor loses its charge in the dark, of PVK:TNF films on different
substrates and charging potentials are depicted in Figure 51.
The variations of spectral sensitivity, i.e., the actual amount of charge dis-
sipated per unit of incident light energy in coulombs/erg, with wavelength
and initial fields are given in Figures 52 and 53. It is clear from Figure 52
130 PENWELL, GANGULY, AND SMITH

40 I Applied voltage

"i
30

25

I
180 200
!Light intensity inpWlcm2

Fig. 46. Photocurrent vs. light intensity for a 1 :1 TPC sample. Reprinted with permission
from the author, SChaiTert (1971). Copyright 1971 by International Business Machines
Corporation.

28
26 - Sarnvle
no. Thickness Substrate
0 I 22.0pm A1 (milled)
24 - 0 I1 36.0pm AI (milled)
A 111 20.4pm A1 (anodized)
22 1V 23.Opm A1 (polished)
V 20.0pm A1 (polished and anodizcd I
20 V IX 45.bpm Nesa
I

IAppIicdfieldinvolts/cmX lom5

Fig. 47. Dark resistivity vs. applied field for several samples of 1 :1 TPC coated on various
substrates. Reprinted with permission from the author, SchaiTert (1971). Copyright 1971 by
International Business Machines Corporation.
POLY(N-VINYLCARBAZOLE) 131

IConcentrationin moles TNF per monomer unit P V C ~


Fig. 48. Dielectric constant as a function of TNF concentration for TNF/PVK photocon-
ductors. Reprinted with permission from the author, Schaffert (1971). Copyright 1971 by
International Business Machines Corporation.

that although the sensitivity of the PVK:TNF system is about one-third that
of selenium in the 400 nm region, the integrated sensitivity over a broad spec-
trum (400-700 nm) is quite comparable to that of selenium or a ZnS:CdS
binder plate.
Sensitivity is directly related to quantum effciency of charge generation.
The dependence of quantum efficiency on wavelength, electric field strength,
and light intensity is shown in Figures 54, 55, and 56, respectively. The re-
ciprocity, defined by the property that the optimum exposure and contrast
potential, for a given input density, d o not change with light intensity, seems
to hold through a wide exposure range (3 psec-1 sec) for a broad-band spec-
trum peaking at about 450 nm. The so-called activation energy, computed
+
from discharge rates at different temperatures (- 8 to 65" C), as a function
of electric field strength is shown in Figure 57. This field dependence of
the activation energy is quite interesting. A near-zero residual with strongly
absorbing continuous exposure and an apparent thickness-independent
photoinduced discharge rate leads to the conclusion that there is no major
deep trapping in the PVK :TNF system.
132 PENWELL, GANGULY, AND SMITH

Fig. 49. Absorption spectrum of the 1:l TPC. Reprinted with permission from the
author, SchafTert (1971). Copyright 1971 by International Business Machines Corporation.

6 0 I I 1 1 I I I I I
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
I Concentration in moles TNF per monomer unit P V C ~
Fig. 50. Charge acceptance as a function of TNF concentration for TNF/PVK coatings
on a clean aluminum substrate (not anodized). Reprinted with permission from the author,
SchafTert (1971). Copyright 1971 by International Business Machines Corporation.
POLY(N-VINYLCARBAZOLE) 133

1, 1'-10.3 ,urn
2.2'- 8.5 !tm

---&--*--

I I I I 1 I I I
0 1 2 3 4 5 6 7 8 9

Time after charging in sec

Fig. 51. Dark decay curves for 1:1 TPC films on several different substrate metals. Nega-
tive (1) and positive charging (1') in aluminum; negative (2) and positive charging (2') on
copper; and negative (3) and positive charging (3') on gold. Reprinted with permission from
the author, Schaifert (1971). Copyright 1971 by International Business Machines Cor-
poration.

300 400 500 700

I Wavelength in nm

Fig. 52. Spectral sensitivity of 1:1 TPC (20 p ) compared to amorphous Se (50 p) and
ZnS:CdS (25 p ; 70% Cds). Curve 1, negative charging; Curve 2, positive charging. R e
printed with permission from the author, Schaffert (1971). Copyright 1971 by International
Business Machines Corporation.
134 PENWELL, GANGULY, AND SMITH

16
14 -
A 5pm fl --
I

--
/

- 1.2 3
o 16pm Left scale,/
/

Fig. 53. Sensitivity vs. initial field strength for several samples of 1 :1 TPC having dif-
ferent thicknesses. Electrostatic contrast of 3.5 X lo-* C/cm2 used for equal charge decay
(left scale); Vo/2used for l / T l , z , Curves 1,2,3 and 4 (right scale). I = 404.5 nm. Reprinted
with permission from the author, Schaffert (1 971). Copyright 1971 by International
Business Machines Corporation.
0.4

0.3 -

g 0.2 -
.8
B
-
.-g
E
0.1

6 0 I I I
300 400 500 600 700
Wavelength in nm
Fig. 54. Quantum efficiency vs. wavelength for the 1 :1 TPC at a constant field strength of
4.0 x lo5 V/cm. Reprinted with permission from the author, Schaffert (1971). Copyright
1971 by International Business Machines Corporation.

Elcctricfield stwngth in voltslcmx lo-'

Fig. 55 Quantum efficiency (QE) vs. field strength. Samples and conditions are the same
as for Figure 54. Solid line: QE for initial discharge vs. initial field; broken line: QE for dis-
charge of 3.5 x C/cma vs. average field strength. Reprinted with permission from the
author, Schaffert (1971). Copyright 1971 by International Business Machines Corporation.
135

ILight intensity in pWlcm2


Fig. 56. Effect of light intensity on quantum efficiencyand photosensitivity of 1 :1 TPC.
1= 404.5 nm; electrostatic contrast = 3.5 x C/cm2; average field strength = 4.5 x
lo5 V/cm. Reprinted with permission from the author, SchafTert (1971). Copyright 1971
by International Business Machines Corporation.

0.20
0.18 -

0.16-
0.14 -

5 0.12 -
.-
C

s 0.10 -
EJ
e 0.08 -
.-
,g 0.06 -
2 0.04 I I I
0 1 2 3

I Electric field strength


Electric field in voltslcmX
strength in voltslcmX lo-’
lo-’

Fig. 57. Activation energy as a function of field strength for a 1 :1 TPC photoconductor.
Reprinted with permission from the author, Schaffert (1971). Copyright 1971 by Inter-
national Business Machines Corporation.

An extensive study of the mobilities of charge carriers in PVK:TNF systems


has been made (Gill, 1972). Field dependences of hole- and electron-drift mo-
bilities for different PVK:TNF ratios are shown in Figures 58 and 59. There
is a common slope for different plots in Figures 58 and 59 with a value of
5.5 x 10-3 (V/cm)-l/2, indicating the same functional form for the field-
136 PENWELL, GANGULY, AND SMITH

Fig. 58. Field dependence of hole-drift mobilities for a range of TNF:PVK molar ratios.
Data taken at T = 24°C. Reprinted with permission from the American Institute of
Physics and the author, Gill (1972).

dependent mobility of electrons and holes. The temperature dependences of


hole- and electron-drift mobilities for 0.2:l molar ratio of TNF:PVK are
shown in Figures 60 and 61. The data extrapolates to a common intersection
at a temperature (TO)of 521°K for hole mobility and 568°K for electron mo-
bility. The field dependence of the activation energy (Eo) calculated from
Figures 58 and 59 is shown in Figure 62. The activation energy is strongly
field dependent and the extrapolated zero field activation energy for 0.2:l =
TNF:PVK system is 0.68 eV for holes and 0.71 eV for electrons. The varia-
tion of hole- and electron-drift mobilities with TNF:PVK molar ratio is shown
in Figure 63. The observed field and temperature dependences of drift mobi-
lity have been fitted by Gill (1972) to a functional form

where
POLY(N-VINYLCARBAZOLE) 137
I I I
I TN F :PVK I

10-6

0
8

N
.
>

-E C
a

lo-’

10-8

Fig. 59. Field dependence of electron-drift mobilities for a range of TNF:PVK molar
ratios. Data taken at T = 24°C. Reprinted with permission from the American Institute of
Physics and the author, Gill (1972).

10-~

-0
%
.
N
>I 1 0 - ~
E
1

0
a
10-5

10-6

10-7
2.0 2.5 3.0 3.5 4.0
1000/T (OK)-’

Fig. 60. Temperature dependence of hole-drift mobility with applied field as a parameter
in a 0.2:1 TNF:PVK film. Reprinted with permission from the American Institute of Phy-
sics and the author, Gill (1972).
138 PENWELL, GANGULY, AND SMITH

1o -~
F = 9.3x lo5 V/cm
TNF:PVK o F = 8 . 0 lo5
~ Vlcm
o F = 5.3x 105 V/cm

1o -~
_-
-
0 F = 2.66x lo5 V/cm
F = 1.33x lo5 Vlcm
F=O

-
g 10-~
.
2.
N

15-
a
10-6

1o-'

10-8
12 1.6 2.0 2.4 2.8 3.2 3.6 4.0
10001T 1 K).'

Fig. 61. Temperature dependence of electron-drift mobility with applied field as a para-
meter in a 0.23 TNF-PVK Elm. Reprinted with permission from the American Institute
of Physics and the author, Gill (1972).

Fig. 62. Activation energies of hole- and electron-drift mobilities a function of applied
field in 0.23 TNF:PVK 5lm. Reprinted with permission from the American Institute of
Physics and the author, Gill (1972).
POLY(N-VINYLCARBAZOLE) 139

TNF:PVK Molar Ratio

Fig. 63. Variation of hole- and electron-drift mobility with TNF:PVK molar ratio. Re-
printed with permission from the American Institute of Physics and the author, Gill (1972).

TABLE 111
Parameters for Exponential Field-Dependent
Mobility in PVK:TNF System
(W. Gill, 1972)

Holes Elect Tons


TNF:PVK -
Eo(eV) T0(OK) ,o(cm
2
/Vsec.)- Eo(eV) T,(OK) ,,o(cm2/Vsec.)

0:1 0.65 660 2x10-2 - -


0.l:l 0.65 625 1x10-2 0.67 595 1.6~1 O-’
0.2:1 0.68 521 ~ . ~ X I O - ~ 0.71 568 3x I f4
0.4:l 0.68 519 4.6xl0-~ 0.68 545 1x10-4
0.6:l 0.65 550 3.8~10’4 - -

The values of the parameter for different compositions and the nature of the
carriers are given in Table 111.
We will discuss briefly the drift mobility of PVK:TNF systems in the light
of possible mechanisms of charge transport in these charge-transfer com-
plexes, and field dependence of photogeneration in terms of different auto-
ionization and field-dependent charge-separation processes in a following
subsection.
140 PENWELL, GANGULY, AND SMITH

E. Photoconductivity of Copolymers and Binder Matrices of N-Vinyl


Carbazole

Very little information is available on the photoconducting properties of


copolymers (random, ordered, or block) of N-vinyl carbazole (NVK). Oka-
mot0 et al. (1973a) have studied photoconductivity of copolymers of NVK
with styrene, vinyl acetate, and n-vinylpyrrolidone. Figure 64 shows the wave-
length dependence of photocurrent (negative charging) for different copoly-
mers of NVK and that of PVK. The photocurrent decreases in all spectral re-
gions for every copolymer. The alternate one-to-one copolymer of NVK and
fumaronitrile shows no photoconductivity at all. The degree of decrease in
photoresponse is apparently independent of the nature of the comonomer
and depends directly on the percentage of comonomer present. An interesting
point to note is that the photosensitivity of PVK doped with 15 % polystyrene
is much higher than the copolytner of NVK:styrene (15 %).

Wavelength (nm)

Fig. 64. Photocurrent of the copolymers of NVK with negative electrode illumination.
(1) PVK, (2) NVK-styrene (15 mol%) copolymer, (3) NVK-vinyl acetate (17 mol%) copol-
ymer, (4) NVK-n-vinylpyrrolidone (20 m o l x ) copolymer. Reprinted with permission
from The Chemical Society of Japan and the authors, Okamoto et al. (1973).

Limburg and Seanor (1975) have studied the photoconducting behavior of


n-vinyl-phthalimide (NVP)/NVK copolymers for different compositions. Fi-
gure 65 gives the action spectrum for positive charging for two different corn-
141

t + V E POTENTIAL

c
15.5

o PURE PVK
o 50150 NVK/NVP
A 75/25 NVK/NVP -
10.0

2Ooo 3000 4000 5Ooo


WAVELENGTH

Fig. 65. Action spectrum of copolymers positive potential. (a) Incident photon flux 1.4 X
1014 photons/cm2at 4700 A. (b) Action spectra normalized to equal incident photons. (c)
Film thickness 14 p, area 1.7 cm2. (d) Semitransparent chromium electrodes. Reprinted
with the permission of the authors, Limburg and Seanor (1975).

positions. The action spectrum for negative charging is similar. As pointed


out by Limburg, the observed enhancement in the photocurrent is primarily
due to the absorption in the charge-transfer band, which extends well into the
visible wavelength. Photoconductivity of the copolymers appears to show a
maximum at 25 mol % of NVP and 75 mol % of NVK. This is ascribed to the
fact that the distribution of donor and acceptor groups is random and with
increasing NVP concentration some of the phthalimide groups are not in-
volved in charge-transfer complexing. Photoconductive behavior of some of
the isostructural modifications of NVP/NVK copolymers, vinylamine/NVK
tetrahalogenated-NVP/NVK and mono- and dinitrovinylaniline/NVK and
copolymers has also been studied by Limburg. Seanor (1975) has studied the
hole mobility in NVP/NVK copolymers and mixtures of PVK and n-ethyl-
phthalimide as functions of field and compositions. The results are shown
142 PENWELL, GANGULY, AND SMITH

w
J
P

20 40 60 80
MOLE oh N-VINYL PHTHALIMIDE

Fig. 66. Hole mobility in PVK: NVP copolymer. Reprinted with the permission of the
author, Seanor (1975).

20 40 60 80
MOLE% N-ETHYL PHTHALIMIDE

Fig. 67. Hole mobility in PVK doped with n-ethylphthalimide.Reprinted with the per-
mission of the author, Seanor (1975).

in Figures 66 and 67 for a field of 80 V/p. The effective hole mobility in the
copolymer system decreases monotonically with an increase in NVP, where-
as the hole mobility seems to increase by doping of N-ethylphthalimide to
PVK and is relatively insensitive to composition as can be seen from Figure
67. Photoconducting properties of graft copolymers of polybutadiene and
NVK and blends of the graft copolymers with PVK have been investigated
(Seanor 1975). The polyblends of PVK showed a composition-dependent pho-
tosensitivity reaching a maximum for pure PVK. The hole mobility is low for
blended systems and selenium sensitization (injection of holes from selenium
POLY(N-VINYLCARBAZOLE) 143

overcoat) seems to be quite difficult. Other binder matrix photoreceptors in-


volving PVK and NVK studied are NVK in Lucite (50 % by weight) (Ellinger,
1964), X-form metal-free phthalocyanine in PVK (Hackett, 1971), and n-
isopropyl carbazole-doped Lexan (Mort, 1976).

F. Photoconductivity of “Pretreated” PVK


In this subsection we shall briefly discuss the effects of impurities, ambient
conditions, radiation (uv light or y rays), heat, and different solvent reten-
tions on the photoconducting behavior of PVK.
Pearson et al. (1974) have carefully studied the effects of impurities (likely
to be present in the commercial PVK) on the absorption spectrum and photo-
induced discharge characteristics of PVK. Figure 68 shows the absorption
spectra of thick films of PVK prepared from purified BASF-LUVICAN, la-
boratory-synthesized PVK, and commercial LUVICAN. The absorption
spectra for purified PVK and synthesized PVK are identical, but commercial
LUVICAN shows a long-wavelength tail. It is concluded that the impurity
producing this tail is not incorporated into the polymer backbone because it
should be removable by repeated reprecipitation and may be a charge-
transfer band of impurity and PVK.

lo5 I I I I - 1 I

-
c

lo2 -

230 270 310 350


WAVELENGTH (nm)

Fig. 68. Absorption spectrum of PVK films on quartz slide. The solid line is common to
synthetic and purified PVK and the dashed line occurs in unpurified commercial samples.
Reprinted with the permission of the authors, Pearson et al. (1974).

The spectral dependence of the initial discharge rates for purified and com-
mercial PVK is shown in Figure 69. At 3500 A, which corresponds to a peak
in the PVK absorption spectrum, the initial discharge rate is approximately
a factor of ten larger in commercial material than in purified commercial PVK
or synthesized PVK. The hole mobility appears to be slightly higher in com-
144 PENWELL, GANGULY, AND SMITH

I
1200 -
I I
-
1000 - -

COMMERCIAL

400 - -
200 - PURIFIED a SYNTHETIC -

300 320 340 360 380


WAVELENGTH (nm)

Fig. 69. Initial xerographic discharge rate vs. wavelength for films of (a) LUVICAN PVK,
(b) synthetic and purified PVK. Reprinted with the permission of the authors, Pearson et
al. (1976).

4- (V/CrnP2 do-2
Fig. 70. The effect of crystallinity of PVK on the hole mobility. Reprinted with the
permission of the authors, Griffiths and Walker (1974).
POLY(N-VINYLCARBAZOLE) 145

mercial PVK than in purified PVK (Sliva, 1974). Air, heat treatment, or
preradiation with uv or y rays seems to increase the photosensitivity of PVK
throughout the spectral region (Okamoto et al., 1973c) in surface-type cell
measurements, but air does not have any drastic effect on the hole mobility
in PVK. X-ray effects on the conductivity of PVK have been studied by
Hughes (1971). There have been some studies on the influence of crystallization
and orientation on the electrical and mechanical properties of polymers (Na-
gro and Herman, 1974; Airapetyants et al., 1964; Prest 1974; Griffiths and
Walker, 1974). The effect of the percentage of crystallinity on the hole mobi-
lity of PVK (Griffiths and Walker, 1974) is shown in Figure 70.
There seems to be no molecular weight dependence of the dark or photo-
conductivity in PVK in the range 3 x 105 to 7 x lo6 (Epping, 1970). HOW-
ever, Tanikawa et al. (1975) did find a molecular weight dependence of pho-
toconductivity in the range 1.2 x 103 to 2.4 x 105. They attribute this poten-
tial discrepancy to a more distinct interruption of the overlap of the w-elec-
trons of adjacent carbonyl groups at the terminal parts of the polymer chains
in the lower molecular weight range.

IX. THEORETICAL STATUS

In this subsection we shall sketch three distinct levels of theoretical under-


standing of the electrical behavior of PVK: first, simulation of the experi-
mental data on photoconduction on the basis of classical electrical field equa-
tions and their solutions under appropriate boundary conditions depending
upon the specific configuration of the photoreceptor. Second, the phenomeno-
logical basis for charge transport and generation in terms of bonds and bands,
and, finally, the stochastic (Markovian and non-Markovian) transport in
disordered solids.

A. Simulation Study of Photoconductivity in PVK


There have been numerous studies to simulate the electrical behavior of
photoconductors in general (Chen and Mort, 1972; Blakney and Grunwald,
1967; Batra et al., 1970a, b, 1971; Batra and Schechtman, 1971; Chen,
1972, 1973; Gill and Batra, 1971; Gill and Kanazawa, 1972; Hermann,
1973; Kanazawa et al., 1972; Kanazawa and Batra, 1972; Many et al., 1962;
Papadakis, 1967; Seki, 1970; Seki and Schechtman, 1972; Seki and Batra,
1970; Seki et al., 1971; Weisz et al., 1968; Ganguly, 1975; Kao and Chen,
1973; Chen and Kao, 1973; Ganguly, 1976; Stockman, 1972) and PVK in
particular (Chen and Mort, 1972; Chen, 1972; Ganguly, 1975; Kao and
Chen, 1973;Chen and Kao, 1973). The basic philosophy of a simulation study
is to simulate the photoconducting behavior of a photoconductor, say PVK,
in terms of various external and internal parameters. The external parameters
are, in general, defined by the experimental conditions (cf. closed circuit or
open circuit, emission-limited or space-charge-limited, etc.) and macroscopic
146 PENWELL, GANGULY, AND SMITH

physical parameters (cf. dimension, dielectric properties, etc.) of the photo-


conductor. The internal parameters are given by the parametrization of many
physical processes pertaining to the particular material such as electric field
and temperature-dependent mobility, shallow and deep trapping, electric-
field-dependent quantum generation of carriers, barrier to injection, lateral
charge conduction, etc. Parametric functional forms of some of these
physical processes relating to PVK and systems involving PVK have been
established experimentally, For example, the temperature and field de-
pendence of carrier mobility (Gill, 1972) are given by Eq. (22) with relevant
parameters tabulated in Table 111. The quantum efficiency (QE) of charge
generation is expressed as

?(v)= 1, v > vq (23b)


where V,, is the potential above which the QE is unity.
The method of simulation of the electrical behavior of a photoconductor
is to incorporate the functional forms of these physical processes into classical
electric field equations (viz., conduction equation, continuity equation, Pois-
son equation, Kirchhoff’s law, etc.) and solve for electric field distribution
inside the photoconductor under appropriate boundary conditions. There is,
so far, no complete treatment of simulation study which includes all plausible
physical processes and solves the field equation exactly. For details of the cal-
culations of the effects of individual processes on photoconductivity the
reader is referred to the original articles. For the most general study of the
case of field-independent mobility, see Batra and Schechtman (1971); for the
power law and exponential law field-dependent case and calculation within
transit time of the fastest carrier, see Gill and Kanazawa (1972); for all plau-
sible cases of field-dependent mobility and calculation in the most general
time domain, see Ganguly (1975); for the power law field-dependent case,
field-dependent quantum generation, and V-mode calculation, see Chen
(1972, 1973); for the bulk generation of carrier, see Kao and Chen (1973)
and Chen and Kao (1973) ;for pigment-dispersedactive matrix PVK, see Gan-
guly (1976); for trapping, see Seki et al. (1970, 1971).

B. Phenomenological Understanding of Photogeneration and


Photoconduction in PVK

I . Photogeneration in PVK
To understand the mechanism of photogeneration in PVK, let us first dis-
cuss the energy-level structure and position of so-called “conduction levels”
in PVK (Stockman, 1972). For an isolated NVK molecule the vacuum ioniza-
POLY(N-VINYLCARBAZOLE) 147

tion energy ZG is about 7.6 eV (Lardon et al., 1967; Lakatos and Mort, 1968 ;
Mort and Lakatos, 1970). In thz solid state, the crystal vacuum level TC is
smaller than ZG by a value of P, corresponding to the polarization energy of
the crystal. The value of P in PVK is about 1.5 eV (Sharp 1967a), leading to
a crystal-state vacuum level of 6.1 eV. Lyons (1957) has loosely defined the
“proper conduction-band energy” as EC = ZG - 2P, which turns out to be
4.6 eV for PVK. The valence band of PVK is considered to be a series of nar-
row (-0.1 eV) bands arising from molecular vibrations.

STI ADI’ STATI A\I) TKA\SII \T PIIOTOI v i w n v

GAS SOLID
0 VACUUY LEVEL (GAS)

I
-VACUUY LEVEL (SOLID)
2

3 - -CONDUCTION IAN0
(eVI
4

?
-- ORWND STATE 01)
VALENCE BAND

Fig. 71. Energy-band diagram for the Cu-PVK interface. The ionization potential (i.e.,
the threshold for photoemission into vacuum) for PVK, ZC, equals ZG - P (ZC = ZG - P)
where ZG is the ionization potential of the PVK molecule and P is the polarization energy
of PVK by a point charge. The notations T and Sn represent triplet and singlet energy
states. Reprinted with permission from the North Holland Publishing Co. and the authors,
Mort and Lakatos (1970).

Mort and Lakatos (1970) have studied the photoinjection of holes from
metals in these valence states of PVK. Schematic band positions of PVK and
hole-injecting metals (Cu, Au) are given in Figure 71. An intrinsic photo-
generation process may be viewed as a combination of two distinct processes.
In the first process, an absorbed photon excites the carrier (electron or hole)
from its ground state to some excited bound state (exciton state). From the
excited bound state the carrier may undergo thermal ionization and/or auto-
ionization (Fano, 1961) into a continuum state, or may decay back to the
ground state by a radiative process, or by a combination of both. There can
be, in general, many processes leading to exciton ionization and/or autoioni-
148 PENWELL, GANGULY, AND SMITH

zation, viz., one- or two-photon processes (Costro and Harding, 1965;


Strome, 1968, respectively), double-exciton interaction processes involving
the singlet-singlet(Braun, 1968) and single-triplet (Fourny et al., 1968) states,
and photoionization of the lowest singlet (Kepter, 1967) and triplet (Holtz-
man et al., 1967) excitons. Once the electrons and holes are in continuous
state they diffuse under the influence of their mutual coulombic attraction and
the applied electric field.
The onset of photoconduction with wavelength does not always coincide
with the absorption edge (Sharp, 1966, 1967a). In n-isopropyl carbazole the
absorption edge is at 3.57 eV, whereas the onset of photoconduction occurs
at 4.7 eV. Crystals of charge-transfer complexes of n-isopropyl carbazole
with picryl chloride showed that the carrier generation still requires 4.7 eV.
The generation efficiency in this system is the same as that of PVK and thus
no carriers are generated by excitation into the charge-transfer absorption
band (Sharp, 1966).
Recently the Onsager mechanism of carrier generation (Onsager, 1938) has
been applied to explain the observed quantum efficiency in PVK and PVK:
TNF systems (Melz, 1972).

2. Photoconduction in PVK
The calculations of the intermolecular overlap energy and the intramole-
cular polaron energy are the determining factors for the mechanism of charge
conduction in organic semiconductors and polymers. A simple foregoing
analysis shows that for band-type conduction, the microscopic mobility must
exceed a value of about 1 cm2/V sec. For a band-type conduction, if z is the
relaxation time of the carrier undergoing multiple scattering which is due to
periodic field, the uncertainty principle dictates

where W is the conduction bandwidth. In a free-electron approximation


(spherical Fermi surface)

W = fi2Ki/2m

where KO = n/a, a being the intermolecular spacing. Now the microscopic


mobility is given by

> - - = - -- 1 cm2/V sec


e e h 4 ea2
p = -z
m m W n R -
for usual values of a in molecular crystals. The microscopic mobility in PVK
is about 2 x cm2/ V sec. The above analysis indicates that it is unlikely
that band-conduction mechanism is valid for PVK. Assuming there is an
intermolecularjump mechanism of conduction in PVK, the mobility can be
calculated from Einstein’s relation
POLY(N-VINYLCARBAZOLE) 149

where D is the diffusion constant. If the probability of transfer of a charge


from one molecule to the neighboring one is Wand the distance between the
interacting molecules is a,the diffusion constant is given by the relation

leading to the expression for microscopic mobility

w
p = (aZe/KIjT>

The probability of transition in unit time induced by electron-phonon inter-


action is given by the well-known expression (Frolich and Sewell 1959)

where

S = (exp(RWo/KBT) - 1)-1

3/ is the intermolecular resonance energy, 10is the Bessel function, y is the


dimensionless electron-phonon interaction, and W Ois the phonon energy.
In the particular case where yS >> 1, the expression for mobility reduces to

P2 wo
yfi
p=--
(: ),"(I? WO)~/~(KBT)~/~ (-m)
The calculation of y is one of the most difficult problems in the jump model
of mobility. y is also intimately connected with the activation energy of the
mobility-a parameter most frequently obtained in experiment.
In the case of PVK, which has a number of polymer chains with conjugated
bonds, the carriers may move along the chain (intrachain jump) and may
jump from one chain to another (interchain jump). If W1 and WZare intra-
chain and interchain jumping probability per unit time, then according to the
Markovian random-walk model (Chandrasekhar, 1943), the mobility is given
by (Kventsel, 1965)

where L is the chain length and N is the number of links in a single chain.
The length- and temperature-dependent mobility are salient features of this
calculation.

C. Continuous-Time Stochastic Conductivity with Absorptive Boundaries


Recently a theory has been developed which takes into account the time
150 PENWELL, GANGULY, AND SMITH

dependence of the stochastic transport processes with absorbing boundaries


(Scher and Lax, 1972, 1973a; Montroll and Scher, 1973; Scher and Montroll,
1975). In this theory one of the points stressed is that the spatial fluctuations
between hopping sites are trivial compared to the fluctuations in waiting
times between hopping events. Central to the time dependence of the hopping
is a proper calculation of y(t), the waiting-time distribution function.
Recent experimental transient photoconductivity in AszSe3 seems to sup-
port many of the predictions of this theory. There are no specific consider-
ations in this theory for polymeric chain structure or activation energy of the
mobility.

X. OTHER VINYL CARBAZOLES

Often in chemistry it is informative, when studying one particular com-


pound, to study similar compounds to determine which molecular features
are most important to the properties of the molecule in question. NVK has
thus been compared with its cogeners, including 3-substituted and 3,6-
disubstitutedderivatives (Inami et al., 1964; Lopatinskii and Zherebtsov, 1965;
Lopatinskii and Shekirev, 1965; Murakami and Morimoto, 1970, 1972;
Pielichowski and Kyziol, 1974; Pielichowski and Puszyniski, 1974; Matsu-
shiro and Oda, 1974; Pielichowski and Daszkiewicz, 1975):

where X = Br, -C1, -I, -N(CH& -NH(CH3), -NOz, -0CH3, -CHzCH3; other
N-alkenyl carbazoles (Heller et al., 1964):

I n = I , 2,3,4

m
and certain N-alkylvinyl carbazoles (Lopatinskii and Sirotkina, 1964a, 1965,
1967; Hyde et al., 1973; Gipstein et al., 1970; Limburg et al., 1975):

I I
R R
lz P
POLY(N-VINYLCARBAZOLE) 151

There have been numerous claims of improved photoconductivity in sub-


stituted N-vinylcarbazoles (Structures I and 11); it is difficult, however, to be
sure of any of these claims.
Many of the N-alkenylcarbazoles (Structure 111) other than N-vinylcarba-
zole were synthesized to study the nature of the Ziegler-Natta polymerization
of vinylcarbazoles. It is interesting in this regard that, while NVK is poly-
merized by most Ziegler-Natta systems, 9-allylcarbazole and 9-d3-butenyl-
carbazole scarcely polymerize. 9-d4-Pentenylcarbazole and 9-d5-hexenylcarb-
azole are polymerized to crystalline polymers by TiC13-EtzAICl(Heller et al.,
1964).
Recently, N-alkyl 2- and 3-vinylcarbazoles (Structures IV and V) have re-
ceived some attention. Limburg et al. (1975) have found that, unlike NVK,
both N-Et-2-VK and N-Et-3-VK are capable of undergoing conventional
anionic polymerization. This discovery with respect to N-Et-3-VK is un-
expected because previous work had indicated the N-Et-3-VK was not
susceptible to anionic polymerization. Poly(N-Et-2-VK) as high in molecular
weight as lo6 an with molecular weight distributions, MWD < 1.5, has been
prepared. In addition, AB and/or ABA-type block copolymers with styrene
and a-Me-styrene, have been prepared. Interestingly, N-Et-2VK is not poly-
merizable by Ziegler-Natta catalyst systems (Hyde et al., 1973). That the
catalyst is deactivated or destroyed by the monomer is evidenced by the im-
mediate color change, purple-brown, which occurs on addition of monomer
to the catalyst.
Limburg and Williams (1973) have been able to correlate the degree of
overlap of carbazole residues (as determined by nmr) in the series PVK, poly-
(N-Et-2-NVK), and poly(N-Et-2-VK) with the mobility of charge carriers in
these polymers. The nmr spectra of poly(N-Et-2-VK) shows the greatest shift
of aromatic protons to a higher field as a result of shielding by adjacent
carbazyl residues. Correspondingly, poly(N-Et-2-VK) exhibits the highest
mobility of charge carriers [1.4 x cm2/V sec, EO(electric field) = 4.0 x
105 V/cm]. PVK’s nmr spectrum reveals considerable shielding of its aromatic
protons and one finds that the mobility of charge carriers in PVK is only
slightly less than that in poly(N-Et-2-VK) (1.4 x 10-7 cm2/V sec, EO = 4.0 x
105 V/cm]. Poly(N-Et-3-VK)’s nmr spectrum displays far less shielding of its
aromatic protons than either PVK or poly(N-Et-2-VK). Correspondingly,
poly(N-Et-3-VK) exhibits the lowest mobility of charge carriers of any
polymer in this series (2 x 10-8 cmZ/V sec, Eo = 4.0 x 105 V/cm) (see Fig.
72). Although poly(N-Et-2-VK) had the highest mobility, it also accumulated
the largest amount of trapped charge. At present it is not possible to identify
the reason for this trapping.
Proton nuclear spin-lattice ( T I )and spin-spin (Tz) relaxation times have
been measured for the series of structurally related polymers, PVK, poly(N-
Et-2-VK) and poly(N-Et-3-VK) (Froix et al., 1976). In conjunction with
dielectric data, the measurements reveal an additional efficient relaxation (6)
in poly(N-Et-2-VK) and poly(N-Et-3-VK) due to ethyl-group rotation. This
relaxation is extremely efficient and leads to a general lowering of the TI
152 PENWELL, GANGULY, AND SMITH

1 SHIELDED I
AROMATIC AROMATIC AL I P HATI C

I I
I I

t I
I I

Fig. 72. 60-MI-h proton nmr spectrum of PVK, P2VK, and P3VK at room temperature
in CDCls. Reprinted with permission from the authors, Limburg and Williams (1973).
Copyright by the American Chemical Society.

values via spin diffusion over those observed for degassed PVK. The y-relax-
ation corresponding to torsional oscillation of the pendant group has a con-
siderably lower activation energy in poly(N-Et-2-VK) and poly(N-Et-3-VK)
than in PVK (4.2, 5.2, and 7.8 kcal/mole, respectively). The T I minima cor-
responding to this relaxation are more intense and occur at lower temperature
in PVK, indicating a narrower distribution of correlation frequencies. These
observations are consistent with a lower degree of steric hinderance in these
polymers than in PVK, which can be easily seen on examinationsof molecular
models; it may be possible to correlate these measurements with the facility
of charge transport in carbazole-containingpolymers.

We would like to express our appreciation to Drs. J. J. O'Malley, C. H. Griffiths, W. W.


Limburg, S. R. Turner, and M. S. Walker for their helpful suggestions during the p r e p
aration of this manuscript.
POLY(N-VINYLCARBAZOLE) 153

REFERENCES
Airapetyants, A. V., Voitenko, R. M., Davydov, B. E., Krentsel, B. A., and Serebryannikova,
V. S.(1964) Vysokomol. Soedin. 6,86.
Akamoto H. and Kuroda H. (1963) J. Chem. Phys. 39, 3364.
Alfrey, T., Jr. and Kapur S. (1949) J. Polym. Sci. 4, 215.
Alfrey, T., Jr., Bohrer, J., and Mark H. (1952a), High Polymers Series, Vol. VIII, Intersci-
ence, New York, p. 40.
Alfrey,T., Jr., Bohrer, J., and Mark, H. (1952b) in High Polymers Series, Vol. W I , Intersci-
ence, New York, pp. 33, 39.
Alfrey, T., Jr., and Magel, B., quoted by R. Hurt (1961) Mukromol. Chem. 47, 143.
Amagi, Y. and Chapiro, A. (1962) J. Chim. Phys. 59, 537.
Asai, M., Kameoka, K., Takeda, Y., and Tazuke, S. (1971) J. Polym. Sci. Polym. Lett.
Ed. 9, 247.
Asia, M. and Tazuke, S. (1973) Macromolecules6, 818.
Asai, M., Tazuke, S.,and Okamura, S. (1974) J. Polym. Sci. Polym. Chem. Ed. 12,45.
Asia, M., Takeda, Y., Tazuke, S., and Okamura, S. (1975) Polym. J. 7, 359.
Ayscough, P. B., Roy, A. K., Groce, R. G., and Munari, S. (1968) J. Polym. Sci. Part
A-I, 6, 1307.
Banford, C. H., Jenkins, A. D., and Ward, J. C. (1960) J. Polym. Sci. 48, 13.
Barkalov, I. M., Goldanskii, V. I., Yenikolopyan, N. S., Terekhova, S. F., and Trofimova,
G. M. (1964) J. Polym. Sci. C4, 909.
Batra, I. P., Kanazawa, K. K., and Seki, H. (1970a) J. Appl. Phys. 41, 3416.
Batra, I. P., Schechtman, B. H., and Kanazawa, K. K. (1970b) Solidstate Commun. 8,1433.
Batra, I . P., Kanazawa, K. K., Schechtman, B. H., and Seki, H. (1971) J. Appl. Phys. 42,
1124.
Batra, I. P. and Schechtman, B. H. (1971) J. Phys. Chem. Solids 32, 769.
Bauer, H. and Klopffer, W. (1970) Chem. Phys. Lett. 7, 137.
Bauser, Von H. and Klopffer, W. (1970) Kolloid 2. Z. Polym. 241, 1062.
Bawn, C. E. H., Fitzsimmons, C., and Ledwith, A. (1964) Proc. Chem. SOC.London,
391.
Bawn, C. E. H., Ledwith, A., and Sambhi, M. (1971) Polymer 12, 209.
Beck, H. and Dorrer, E. (1940) U. S. Pat. 2,215,573, to General Aniline and Film Corp.
Bergfjord, J. A. and Penwell, R. C. (1975) private communication.
Bevington, J. C. and Dyball, C. J. (1975) J. Chem. SOC.Furuduy Trus. 71,2226.
Billmeyer, F. W., Jr. (1971) Textbook of PoIymer Science, 2nd ed., Wiley, New York.
Biswas M. and Chakravorty, D. (1970) Bull. Chem SOC.Jpn. 43, 1904.
Biswas, M. and Chakravorty, D. (1973) J. Polym. Sci. Polym. Chem. Ed. 11, 7.
Biswas, M. and Chakravorty, D. (1974) J. Polym. Sci. PoIym. Chem. M.12, 1337.
Biswas, M. and Kamannarayana, P. (1975) J. Polym. Sci. 13, 2035.
Biswas, M. and Ghosal, S. (1966) Chem. Znd. 41, 17.
Biswas, M. and Kar, I. (1967) Indian J. Chem. 5, 119.
Blakney, R. M. and Grunwald, H. P. (1967) Phys. Rev. 159, 658.
Bourdon, J. and Schnuriger, B. (1967) Physics and Chemistry of the Organic Solid State,
Vol. 111, Wiley-Interscience, New York, p. 59.
Bowyer, P. M., Ledwith, A., and Sherrington, D. C. (1971) Polymer 12,509.
Brandrup, J . and Immergut, E. H. (1974) Polymer Hundbook, 2nd ed., WiIey, New York.
Braun, C. L. (1968) Phys. Rev. Lett. 21, 215.
Bree, A. and Zwarich, R. (1968) J. Chem. Phys. 49, 3344.
Breitenbach, J. W. and Sma, C. (1962) Pure Appl. Chem. 4,245.
Breitenbach, J. W., Olaj, 0. F., and Wehrman, F. (1964) Monutsh. Chem. 95, 1007.
Breitenbach, J. W. and Olaj, 0. F. (1964) J. Polym. Sci. Polym. Lett. Ed. 2,685.
British Pat. 739,438 (1955) to Badische Anilin-Soda Fabrik.
British Pat. 914,418 (1963) to Montecatini Societa Generale per L'lndustriu Mineravia E
Chimicu, Chem. Abstracts 58, 9253 (1963).
154 PENWELL, GANGULY, AND SMITH

Busse, W. F., Lambert, J. M., McKinley, C., and Davidson, H. R. (1948) Ind. Eng. Chem.
40, 227 1.
Busse, W. F. and Lambert, J. M., U. S. Pat. 2,618,020 (1952) to General Aniline and Film
Corp.
Burkhart, R. D. (1976) Macromolecules 9, 234.
Carnes, J. E. and Warter, P. J., Jr. (1972) Phys. Rev. B4,1557.
Cassiers, P. M. and Hart, R. M. (1971) Can. Pat. 887,578 to Gevaert Photo-Production
N. V., Mortsel-Antwerp, Belgium.
Castro, G. and Hornig, J. F. (1965) J. Chem. Phys. 42, 1459.
Castro, G. (1971) ZBM J. Res. Dev. 15,27.
Chandrasekhar, S . (1943) Rev. Mod. Phys. 15,2.
Chapiro, A. and Hardy, G. (1962) J. Chim. Phys. 59,993.
Chapiro, A., U. S. At. Energy Comm. TID-7643,136-149 (1972), Chem. Abstracts 58,6933
(1963).
Chapiro, A. (1963) Magy. Kem. Lupja 18 152; Chem. Abstracts 59, 2312 (1963).
Chapiro, A. and Mankowski, Z. (1973) C. R. Acud. Sci. Ser. C. 277,291.
Chen, I. and Mort, J. (1972) J. Appl. Phys. 43, 1164.
Chen, I. (1972) J. Appl. Phys. 43,1137; see also (1973) J. Appl. Phys. 44,921.
Chen, I. and Kao, C. (1973) J. Appl. Phys. 44, 2718.
Chernobai, A. V., Tirakyants, Zh. S., and Delyatitskaya, R. Ya. (1967) Vysokomol. Soedin.
9,664.
Chu, J. Y. C. and Stolka, M. (1975) J. Polym. Sci. Polym. Chem. Ed. 13,2867.
Ciba (A. R. L.) Ltd., Fr. Pat. 1,397,538, (1965); Chem. Abstructs 63,7131 (1965).
Cornish, E. H. (1963) Plustics 28, 61.
Crellin, R. A. and Ledwith, A. (1975) Macromolecules 8, 93.
Crystal, R. G. (1971) Macromolecules 4, 379.
Dall’Asta, G. and Casale, A. (1966) Chem. Abstracts 65, 3968.
Davidge, H. (1957) J. Appl. Chem. 9, 553.
Davidge, H. (1959) J. Appl. Chem 9, 241.
Davies, D. H., Phillips, D. C., and Smith, J. D. B. (1973) J. Org. Chem. 38, 2562.
Denbigh, K. G. (1940) Trans. Furuduy SOC.36,936.
Dev, S . B., Lockhead, R. Y., and North, A. M. (1970) Discuss. Furuduy Soc. 49, 244.
Distillers Co. Ltd., Neth. Appl. 6,612,244 (1967); Chem. Abstracts 67, 64851 (1967).
Dumont, E. and Reinhardt, H. (1957) Ger. Put. 1,001,488.
Ellinger, L. P. (1964) Polymer 5 , 559.
Ellinger, L. P. (1965) J. Appl. Polym. Sci. 9, 3939.
Ellinger, L. P. (1965) Polymer 6, 549.
Ellinger, L. P. (1966) J. Appl. Polym. Sci. 10, 551.
Ellinger, L. P. (1966) J. Appl. Polym. Sci. 10, 575.
Ellinger, L. P. (1969) Polymer 10, 531.
R. H. Epping, Inaugural-Dissertation zur Erlangung des Doktogrades; Naturwissen-
schaften. Fakultat ; Universitat Miinchen (1970), also : in Current Problems in Electro-
photography, W. F. Berg and K. Hanffe, Eds., deGruyter, Berlin, 1972, p. 215.
Erhardt, P. and Richards, W. C. (1975) unpublished data.
Ershov, Yu A., Kujina, S. I., and Neiman, M. B. (1969) Russ. Chem. Rev. 38, 147.
Fano, U. (1961) Phys. Rev. 124, 1866.
Ferry, J. D. (1971) Viscoelustic Properties of Polymers, Wiley, New York.
Fikentscher, H. and Fricker, R. (1955) Ger. Pat. 931,731, Badische Anilin und Soda Fabrik.
Flory, P. J. (1953) Principles of Polymer Chemistry, Cornell Univ. Press, Ithaca, New York.
Foden, E., Morrow, D. R., and Sauer, J. A. (1972) J. Appl. Polym. Sci. 16,519.
Foumy, J., Delacote, G., and Schott, M. (1968) Phys. Rev. Lett. 21, 1085.
Fritz, E. G. (1964) NARF-63-14T, MR-N-307.
Froix, M., Williams, D. J., and Goedde, A. 0. (1975) J . Appl. Phys. 46,4166.
Froix, M. F., Williams, D. J., and Goedde, A. 0. (1976) Macromolecules 9, 81.
Frolich, H. and Sewell, G. L. (1959) Proc. Phys. SOC.74, 643.
POLY(N-VINYLCARBAZOLE) 155

Galdecki, Z., Karolak, J., Pekala, W., and Kroh, J. (1967) Bull. Acud. Pol. Sci. Ser. Sci.
Chim. 15, 209.
Ganguly, B, N. (1975) Phys. Rev. 12, 1275.
Ganguly, B. N., J. Appl. Phys., in press.
Gibbons, D. J. and Spear, W. E. (1966) J. Phys. Chem. Solids 27, 1917.
Gill, W. D. and Batra, I. P. (1971) J . Appl. Phys. 42, 2067.
Gill, W. D. (1972) J. Appl. Phys. 43, 5033.
Gill, W. D. and Kanazawa, K. K. (1972) J . Appl. Phys. 43,529.
Gipstein, E., Hewett, W. A., and Need, 0. U. (1970) J. Polym. Sci. Part A 8, 3285.
Golden, J. H., Hammant, B. L., and Hazell, E. A. (1967) J. Appl. Polym. Sci. 11, 1571.
Grassie, N. and Weir, N. A. (1965a-d) J. Appl. Polym. Sci. 9,963, 975, 987,999.
Grassie, N. (1972) Pure Appl. Chem. 30, 119.
Griffiths,C. H. and Van Laeker, A. (1976)J. Polym. Sci. Polym. Phys. Ed. 14,1433.
Griffiths, C. H. and Walker, M. S. (1974) 13thIUPACInt. Symp.Macromol.,Madrid, Spain.
Griffiths, C. H. (1975) J. Polym. Sci. Polym. Phys. Ed. 13, 1167.
Guillet, J. E. (1972) Pure Appl. Chem. 30, 135.
Hackett, C. F. (1971) J. Chem. Phys. 55, 3178.
Hardy, G., Cser, F., Kovacs, G., Fedorova, N., and Samay, G. (1973) Actu Chim. 79,143.
Hayashi, Y., Kuroda, M., and Inami, A. (1966) Bull. Chem. SOC.Jpn. 39, 1660.
Heller, J., Tieszen, D. O., and Parkinson, D. B. (1963) J. Polym. Sci. Part A 1, 125.
Heller, J., Lyman, D. J., and Hewett, W. A. (1964) Mukromol. Chem. 73,48.
Hermann, A. M. (1973) J. Appl. Phys. 44,926.
Hernandez, C. and Gandini, A. (1973) Rev. Cenic Cienc. Fis. 4,49.
Hoegl, H. (1965) J. Phys. Chem. 69, 755.
Hoegl, H., Barchietto, G., and Tar, D. (1972) Photochem. Phofobiol. 16, 335.
Holtzman, P., Morris, R., Jarnagin, R. C., and Silver, M. (1967) Phys. Rev. Lett. 19, 506.
Hughes, J. and North, A. M. (1964) Proc. Chem. SOC.,404.
Hughes, J. and North, A. M. (1966) Trans. Furuduy SOC.62, 1866.
Hughes, R. C. (1971) Chem. Phys. Lett. 8,403.
Hughes, R. C (1972) Appl. Phys. Lett. 21, 196.
Hurt, R. (1961) Mukromol. Chem. 47, 143.
Hyde, P., Kricka, L. J., and Ledwith, A. (1973) Polymer 14,124.
Hyde, P. and Ledwith, A. (1974) in Molecular Complexes, Vol. 11, Foster, R., Ed., Paul,
London, p. 189.
Ikeda, M., Morimoto, K., Murakami, Y., and Sato, H. (1969) Jpn. J. Appl. Phys. 8,759.
Ikeda, M., Morimoto, K., Murakami, Y., and Sato, H. (1970) Jpn. J. Appl. Phys. 9,931.
Inami, A., Morimoto, K., and Murakami, Y. (1964) Nippon Kaguku Zusshi 85,880; Chern.
Abstracts , 62, 14612C (1965).
Iyer, P. N. S., North, A., Petrik, R. A., and Steinhauer, D. B. (1975) Polymer 16, 797.
Jellineck, H. G. (1967) Mod. Plust 44, 160.
Johnson, G. E. (1974) J. Chem. Phys. 61, 3002.
Jones, R. G., Catterall, E., Bilson, R. T., and Booth, R. G. (1972) J. Chem. SOC.Chem.
Commun. 1, 22.
Joshi, R. M. (1962) Mukromol. Chem. 55, 33.
Kanazawa, K. K., Batra, I. P., and Wintle, H. J. (1972) J. Appl. Phys. 43, 719.
Kanazawa, K. K. and Batra, I. P. (1972) J. Appl. Phys. 43, 1845.
Kao, C. and Chen, I. (1973) J. Appl. Phys. 44,2708,2718.
Kauffmann, H. F., Breitenbach, J. W., and Olaj, 0. F. (1973) J. Polym. Sci. Polym.
Chem. Ed. 11, 737.
Kepter, R. G. (1967) Phys. Rev. Lett. 18, 951.
Kikuchi, Y. and Fukuda, H. (1974) Nippon Kuguku Kuishi 1, 200.
Kikuchi, Y. and Ueyama, T. (1974) Nippon Kuguku Kaishi5,981.
Kimura, A., Yoshimoto, S., Akana, Y., Hirata, H., Kusabayashi, S., Mikawa, H., and
Kasai, N. (1970) J. Polym. Sci. Part A 8, 643.
Klopffer, W. and Willicks, W. (1968) Mucromol. Chem. 115,156.
156 PENWELL, GANGULY, AND SMITH

Klopffer, W. (1969a) J. Chem. Phys. 50, 2337.


Klopffer, W. (1969b) Chem. Phys. Lett. 4, 193.
Klopffer, W. and Liptay, W. (1970) 2. Naturforsch. 25a, 1091.
Klopffer, W. and Fischer, D. (1973) J. Polym. Sci. Polym. Symp. 40,43.
Kroh, J. and Pekala, W. (1964) Bull. Acad. Pol. Sci. Ser. Sci. Chim. 12,419; Chem. Abstracts
62, 1240 (1965).
Kroh, J. and Pekala, W. (1966) Bull. Acad. Pol. Sci. Ser. Sci. Chim. 14,55; Chem. Abstracts
64,19795 (1966).
Kuroda, H., Hasegawa, K., and Schneider, W. G. (1964) Can. J. Chem. 42, 1084.
Kuwahara, N., Higashida, S., Nakato, M., and Kaneko, M. (1969) J. Polym. Sci. Part A
7, 285.
Kventsel, G. F. (1965) Zh. Eksp. Teor. Khim. 1, 826.
Lakatos, A. I. and Mort, J. (1968) Phys. Rev. Lett. 21, 1444.
Lanker, W. (1972) in Current Problems in Electrophotography,Berg, W. F.and Hanffe, K.
Eds., deGruyter, Berlin.
Lardon, M., Lell-Doller, E., and Weigl, J. W. (1967) Mol. Cryst. 2, 241.
Ledwith, A. and Shemngton, D. C. (1972) Macromol. Synth. 4, 184.
Ledwith, A. (1967) J. Appl. Chem. 17, 344.
Ledwith, A. (1975) private communication.
Ledwith, A. (1972) Acc. Chem. Res. 5, 133.
Legrand, D. G. (1969) J. Appl. Polym. Sci. 13,2129.
Lell, E. and Weigl, J. W. (1964) Abstr. Znt. Con$ on Photosensitization in Solids, Chicago.
Lenz, R. W. (1967) Organic Chemistry of Synthetic High Polymers, Interscience, New York.
Limburg, W. W. and Seanor, D. A. (1975) U. S. Pat. 3,877,936, to Xerox Corp.
Limburg, W. W., Yanus, J. F., Williams, D. J., Goedde, A. O., and Pearson, J. M. (1975), J.
Polym. Sci. Polym. Chem. Ed. 13, 11 33.
Limburg, W. W. and Williams, D. J. (1973) Macromolecules6,787.
Lopatinskii, V . P., Sirotkina, E. E., and Anasova, M. M. (1964a) Tr. Tomsk. Gos. Univ.,
Ser. Khim. 170,49; Chem. Abstracts 63, 563 (1965).
Lopatinskii, V. P. and Sirotkina, E. E. (1964b) Melody Polucheniya Khim. Reakitvov
Preparatov, 11, 40; Chem. Abstracts 63, 2203 (1966).
Lopatinskii, V. P. and Zherebtsov, I. P. (1965) Zsv. Tomsk. Politekh. Znst. 136, 23; Chem.
Abstracts 65, 18550 (1966).
Lopatinskii, V. P. and Shekirev, Yu. P. (1965) Polytechnical Institute, Tomsk, USSR Pat.
193, 770 (Sept. 7,1965); Chem. Abstracts 69,2060 (1966).
Lopatinskii, V. P. and Sirotkina, E. E. (1965) Zzv. Tomsk. Politekh. Znsr. 136, 26; Chem.
Abstracts 65, 16930 (1966).
Lopatinskii, V. P., Sirotkina, E. E., and Pakalskaya, S. D. (1967) Zzv. Tomsk. Politekh. Znsr.
148,70; Chem. Abstracts 70, 87441 (1969).
Lyons, L. E. (1957) J. Chem. SOC.Pt. 4, 5001.
Mankowski, Z. (1973) J. Chim. Phys. Physicochim.Biol. 70,1299.
Many, A. and Rakavy, G. (1962) Phys. Rev. 126,1980.
Many, A., Weisz, S. Z., and Simhony, M. (1962) Phys. Rev. 126,1989.
Many, A. (1965) J. Phys. Chem. Solids 26,515.
Mark, P. and Helfrich, W. (1961) J. Appl. Phys. 33,205.
Matsuda, T., Higashimura, T., and Okamura, S. (1967) Kobunshi Kagaku 24, 165.
Matsuda, T., Higashimura, T., and Okamura, S. (1968) J. Macromol. Sci. Chem. A2, 43.
Matsushiro, K. and Oda, T. (1974) Jpn. Pat. 74, 09466, Nippon Synthetic Chemical
Industry Co., Ltd.
Meier, H. (1963) Die Photochemieder Organischen Farbstofe, Springer-Verlag, Berlin, p.
158 [(1968) Spectral Sensitization, Focal Press, London.]
Meier, H. (1965) Angew. Chem. 77, 663.
Meier, H., Albrecht, W., and Tschirwitz, U. (1969) Ber. Bunsenges. Phys. Chem. 73, 795.
Meier, H., Albrecht, W., and Tschirwitz, U. (1972a) Photochem. Photobiol. 16, 353.
Meier, H., Albrecht, W., and Tschirwitz, U. (1972b) in Current Problems in Electrophoto-
POLY(N-VINYLCARBAZOLE) 157

graphy, Berg, W. F. and Hanffe, K. Eds., deGruyter, Berlin, p. 163.


Melz, P. J. (1972) J. Chem. Phys. 57, 1694.
Meyers, R. A. (1966) Western Reg. Meet. of the ACS, San Francisco.
Meyers, R. A. and Christman, E. M. (1968) J. Polym. Sci. Part A 6, 945.
Miller, S. A. and Davidge, H. (1958) U. S. Pat. 2,830,059, to British Oxygen Co., Ltd.
Mininni, R. M., Moore, R. S., Flick, J. R., and Petrie, S. E. B. (1973) J. Macromol. Sci.
Phys. B8,343.
Miyama, H., Tsuji, J., Morikawa, M., Kamachi, M., and Nod, T. (1967) Japan Pat. 16,049,
to Toyo Rayon, Ltd.
Montillier, J-P. (1974) U. S. Pat. 3,811,879.
Montroll, E. W. and Scher, H. (1973) J. Stat. Phys. 9, 101.
Mort, J. and Lakatos, A. I. (1970) J. Non-Cryst. Solids 4, 117.
Mort, J. (1972) Phys. Rev. B5, 3329.
Mort, J. and Neilsen, P. (1972) Phys. Rev. B5, 3336.
Mort, J., Chen, I., Emerald, R. L. and Sharp, J. H. (1972) J. Appl. Phys. 43,2285.
Mort, J. and Emerald, R. L. (1974) J. Appl. Phys. 45, 175.
Mort, J., Pfister, G., and Grammatica, S. J. (1976) Bull. Am. Phys. SOC.21, 315.
Murakami, Y. and Morimoto, K. (1970) U. S. Pat. 3,526,502,to Matsashita Electric
Industrial Co., Ltd.
Murakami, Y. and Morimoto, K. (1972) Jupun Kokui (JKXXAF) 74 11876 (16E333).
Naghizadeh, J. and Springer, J. (1967) Kolloid Z . Z . Polym. 215, 21.
Nagao, M. and Hermann, A. M. (1974) J. Polym. Sci. Polym. Lett. Ed. 12, 69.
Nakamura, T., Soma, M., Onishi, T., and Tamura, K. (1970) Makromol. Chem. 135, 341.
Natsuume, T., Nishimura, M., Fujimatsu, M., Shimizu, M., Shirota, Y., Hirata, H., Ku-
sabayashi, S., and Mikawa, H. (1970) Polym. J. Jpn. 1, 181.
Nelson, R. C. (1956) J. Opt. SOC.Am. 46,13.
Nishii, M., Tsuji, K., Takakura, K., Hayashi, K., and Okamura, S. (1964-65) Ann. Rep.
Jpn. Assoc. Rad. Res. Polym. 6, 181 .
Nishii, M., Tsuji, K., Hayashi, K., Okamura, S., and Takakura, K. (1966) Kobunshi Kagaku
23,445.
North, A. M. and Phillips, P. J. (1968) Chem. Commun. 21, 1340.
North, A. M. and Whitelock, K. E. (1968) Polymer 9, 590.
Ohno, A., Koti, N., and Kawase, N. (1972) Polym. Lett. 10, 133.
Oka, S., Mori, T., Kusabayashi, S., Yamamoto, Y., Ishiguro, M., and Mikawa, H. (1964a)
Electrophoiography Jpn. 5, 52.
Oka, S., Mori, T., Kusabayashi, S., Taniguchi, A., Yamamoto, Y., Ishiguro, M., and
Mikawa, H. (1964b) Electrophotogruphy Jpn. 5,77.
Okamoto, K., Hasegawa, Y., Kusabayashi, J., and Mikawa, H. (1968) Bull. Chem. SOC.
Jpn. 41, 2563.
Okamoto, K., Kusabayashi, S., and Mikawa, H. (1973a) Bull. Chem. SOC. Jpn. 46,2613.
Okamoto, K., Kato, K., Murao, K., Kusabayashi, S., and Mikawa, H. (1973b) Bull. Chem.
SOC.Jpn. 46,2883.
Okamoto, K., Kusabayashi, S., and Mikawa, H. (1973~)Bull. Chem. SOC.Jpn. 46, 1948.
Okamura, S., Hayashi, K., and Kitanishi, Y. (1962) J. Polym. Sci. 58,925.
Okamura, S., Higashimura, T., and Matsuda, T. (1965) Kobunshi Kugaku 22, 80.
Okamura, S., Higashimura, T., and Matsuda, T. (1966) Kobunshi Kugaku 23, 269.
Olaj, 0. F., Breitenbach, J. W., and Kauffman, H. F. (1971) J. Polym. Sci. Polym. Lett.
Ed. 9, 877.
Onsager, L. (1938) Phys. Rev. 54, 554.
Orinik, M. T. (1948) U. S. Pat. 2,453,097 to General Aniline and Film Corp.
Pac, J. and Plesch, P. H. (1967) Polymer 8,237.
Pai, D. (1970) J. Chem. Phys. 52,2286.
Papadakis, A. C. (1967) J. Phys. Chem. Solids 28, 641.
Pearson, J. M., Yanus, J. F., Williams, D. J., and Goedde, A. 0. (1974) private communi-
cation.
158 PENWELL, GANGULY, AND SMITH

Penwell, R. C. and Prest, W. (1976) Am. Chem. Soc. Polym. Prepr. 17,931.
Peredereeva, S . I. Orlov, I. G., Zaikov, G. E., and Cherkashin, M. I. (1970) Dokl. Akad.
Nauk. SSR 195, 1150.
Phillips, D. C., Davies, D. H., and Smith, J. D. B. (1972) Macromolecules 5,674.
Phillips, D. C., Davies, D. H., and Smith, J. D. B. (1973) Makromol. Chem. 169, 177.
Pielichowski, J. (1973) J. Polym. Sci. Polym. Symp. 42,451.
Pielichowski, J. and Kyziol, J. (1974) J. Polym. Sci.Polym. Lett. Ed. 12,257.
Pielichowski, J. and Kyziol, J. (1974) Monatsschr. Chem. 105, 1306.
Pielichowski, J. and Puszyniski, A. (1974) Monatsschr. Chem. 105, 772.
Pielichowski, J. and Daszkiewicz, Z. (1975) Rocz. Chem. 49, 1581.
Pochan, J. M., Hinman, D. F., and Nash, R. (1975) J. Appl. Phys. 46 , 4115.
Prest, W. M. (1974) private communication.
Privalko, V. P. (1973) Macromolecules 6, 11 1.
Queener, C. A. (1971) Photogr. Sci. Eng. 15,423.
Rabinowitz, S . and Beardmore, P. (1972) CRC Crit. Rev. Macromol. Sci.
Regensburger, P. J. (1968) Photochem. Photobiol. 8, 429.
Repjx, W., Keyssner, E., and Dorrer, E. (1935) Fr. Pat. 792,820.
Restaino, A. J., Mesrobian, R. B., Morawetz, H., Ballantine, D. S., Dienes, G. J., and
Metz, D. J. (1957) J. Am. Chem. Soc. 78, 2939.
Barrales-Rienda, J. M., Gondlez-Ramos, J., and Dabrio, M. V. (1976) Angew. Makro.
Chem. 43, 105.
Sato, H. and Ikeda, M. (1972) J. Appl. Phys. 43,4108.
Schaffert, R. M. (1971) ZBM J. Res. Dev. 15, 75.
Scher, H. and Lax, M. (1972) J. Non-Cryst. Solids 8-10, 497.
Scher, H. and Lax, M. (1973a) Phys. Rev. B7,4491.
Scher, H. and Lax, M. (1973b) Phys. Rev. B7, 4502.
Scher, H. and Montroll, E. W. (1975) Phys. Rev. B12, 2455.
Schildknecht, C. E., Zoss, A. O., and Grosser, F. (1949) Znd. Eng. Chem. 41, 2891.
Scott, H., Miller, G. A., and Labes, M. M. (1963) Tetrahedron Lett. 17, 1073.
Scott, H. and Labes, M. M. (1963) J. Polym. Sci. Polym. Lett. Ed. 1, 413.
Scott, H., Konen, P., and Labes, M. (1964) J. Polym. Sci. Polym. Lett. Ed. 2, 689.
Seanor, D. A. (1975) private communication.
Seki, H. (1970) Phys. Rev. B2,4877.
Seki, H. and Batra, I. P. (1970), J. Appl. Phys. 41, 3409.
Seki, H. and Batra, I. P. (1971) J. Appl. Phys. 42, 2407.
Seki, H., Batra, I. P., Gill, W. D., Kanazawa, K. K., and Schechtman, B. H. (1971) ZBM
J. Res. Dev. 15, 213.
Seki, H. and Schechtman, B. H. (1972) J. Appl. Phys. 43, 523.
Sharp, J. H. (1966) J. Phys. Chem. 70, 584.
Sharp, J. H. (1967a) J . Phys. Chem. 71, 2587.
Sharp, J. H. (196713) Photogr. Sci. Eng. 11, 69.
Shattuck, M. D. and Vahtra, U. (1969) U. S. Pat. 3,484,237.
Shimada, M. and Tanaka, T. (1971) Jpn. Pat. 7,354,190 [26(3)B324].
Shirota, Y., Kawai, K., Yamamoto, N., Tada, K., Shida, T., Mikawa, H., and Tsubomura,
H. (1972) Bull. Chem. Soc. Jpn. 45,2683.
Shirota, Y., Matsumoto, A., and Mikawa, H. (1972) Polym. J. 3, 643.
Shirota, Y., Yokoyama, M., and Mikawa, H. (1973) Am. Chem. Sci. Polym. Prepr. 14,
13.
Sitaramaiah, G . and Jacobs, D. (1970) Polymer 11, 165.
Sliva, P. 0. (1974) private communication.
Smith, M. and Weigl, J. W. (1965) in Xerography and Related Processes, Dessauer, J. H.
and Clark, H. E., Eds., Focal Press, New York, pp. 169-199.
Solomon, 0. F., Dimonie, M., Ambrozh, K., and Tomesku, M. (1961) J. Polym. Sci. 52,205.
Solomon, 0. F., Cinta, I. Z., and Cobianu, N. (1964) J. Polym. Sci. Polym. Lett. Ed. 2,
311.
POLY(N-VINYLCARBAZOLE) 159

Solomon, 0. F., Cobianu, N., and Cinta, I. Z. (1965) Inst. Bull. Politech. Bucuresti27,65;
Chem.Abstracts 64, 3693 (1966).
Spear,W. E. (1969) J. Non-Cryst. Solids 1, 197.
Stastny, F. and Gerlich, H. (1955) Ger. Pat. 934,692,to Bad. Anilin and Soda-Fabrik Akt.-
Ges.
Stastny, F. and Buchholz, K. (1965) Ger. Pat. 951,229 to Bad. Anilin and Soda-Fabrik
Akt.-Ges.
Stein, R. S. (1958) J. Polym. Sci. 31, 327.
Stein, R. S. (1975) Am. Chem. Soc Polym. Prepr 16,387.
Stein, R. S. and Su, T. K. (1975) Bull. Am. Phys. SOC.20, 340.
Stevenson, D. P. and Coppinger, G. M. (1962) J. Am. Chem. SOC.84,149.
Stockman, D. L. (1972) Current Problems in Electrophotogruphy,Berg, W. F. and Hanffe,
K., Eds., deGruyter, Berlin, p. 194.
Stolka, M. and Turner, S. R. J . Polym. Sci. Polym. Chem. Ed., to be published.
Stokla, M. (1976) U. S. Pat. 3,994,994 to Xerox Corp.
Strome, F. C. (1968) Phys. Rev. Lett. 20, 3.
Su, T. K. (1975) Ph. D. thesis, Univ. of Mass.
Sutyagin, V. M., Lopatinskii, V. P., Sirotkina, E. E., and Berzin, V. I. (1974) Vysokomol.
Soedin. Ser. B 16,452.
Szymanski, S. and Labes, M. M. (1969) J. Chem. Phys. 50,3568.
Tabak, M. D. (1967) Trans. AIME 239, 330.
Tada, K., Shirota, Y., and Mikawa, H. (1973) Macromolecules6,9.
Tada, K., Shirota, Y., and Mikawa, H. (1972) J. PolymSci. Polym. Lett. Ed. 10, 691.
Tagawa, S., Tabata, Y., Arai, S., and Imamura, M. (1974) J. Polym. Sci. Polym. Lett. Ed.
12, 545.
Takeda, Y.,Asai, M., and Tazuke, S. (1975) Polym. J. 7, 366.
Tanikawa, K., Enomoto, T., Hatano, M., Motegi, K., and Okuno, Z. (1975) Macromol.
Chem. 176,3025.
Tazuke, S., Nakagawa, K., and Okamura, S. (1965) J. Polym. Sci Polym. Lett. Ed. 3, 923.
Tazuke, S., Asai, M., Ikeda, S., and Okamura, S. (1967) J. Polym. Sci. Polym. Lett. Ed.
5,453.
Tazuke, S., Asai, M., and Okamura, S. (1968) J. Polym. Sci. Part A , 6, 1809.
Tazuke, S. and Okamura, S. (1968a) J. Polym. Sci. Part A 6,2907.
Tazuke, S . and Okamura, S. (1968b) J. Polym. Sci. Polym. Lett. Ed. 6, 173.
Tazuke, S. (1970) J. Chem: SOC.Chem. Commun. 20, 1277.
Tazuke, S. (1970) J. Phys. Chem. 74,2390.
Tazuke, S. (1973) Pure Appl. Chem. 34, 329.
Treloar, L. R. G. (1958) 7he Physics of Rubber Elasticity, Oxford Univ. Press, Cambridge.
Tsuchihashi, N., Enomoto, T., Tanikawa, K., Tajiri, A., and Hatano, M. (1975) Makromol.
Chem. 176,2833.
Tsuji, K., Takakura, K., Wishii, M., Hayashi, K., and Okamura, S. (1964-65) Ann. Rep.
Jpn. Assoc. Rad. Res. Polym. 6,179.
Tsuji, K . (1973) Adv. Polym. Sci. 12,131.
Turchi, G., Matera, F., and Magagnini, P. L. (1973) Mukromol. Chem. 170, 75.
Turner, S. R. (1975) private communication.
Boros-Tyevi, B. (1969) Magy. Kem. Foly. 75,87; Chem.Abstracts 70,88284 (1969).
Ueberreiter, K. and Springer, J. (1963) 2. Phys. Chem. 36,299.
Ueberreiter, K . and Bruns, W. (1964) Ber. Bunsenges Phys. Chem. 68, 541.
Ushakov, S. N. and Nikolaev, A. F. (1956) Izv. Akad. Nauk. USSR Otd. Khim. Nauk. 83;
Chem.Abstracts 50, 13867 (1956).
Ushakov, S. N. and Nikolaev, A. F. (1956a) Bull. Acad. Sci. USSR Div. Chem. Sci. (Engl.
transl.) 79, 1.
Ushakov, S. N. and Nikolaev, A. F. (195613) Bull. Acad. Sci. USSR Div. Chem. Sci. (Engl.
transl.) 79, 217.
Van Krevelin, D. W. and Hoftyzer, P. J. (1967) J. Appl. Polym. Sci. 11, 1409.
160 PENWELL, GANGULY, AND SMITH

Wagner, W. J. and Gasner, E. L. (1970) Photogr. Sci. Eng. 14,205.


Weigl, J. W. (1963) Photographic Science Symposium 1961, Focal Press, New York, pp.
345-370.
Weigl, J. W. (1972) Photochem. Photobiol. 16,291.
Weisz, S . W., Cobas, A., Trester, S., and Many, A. (1968) J. Appl. Phys. 39, 2296.
Wignall, G. D. and Longman, G. W. (1974) Am Chem. SOC.Polym. Prepr. 15,18.
Williams, J. K., Wiley, D. W., and Muchusick, B. C. (1962) J. Am. Chem. SOC.84, 2216.
Williams, D. J. (1970) Macromolecules 3, 602.
Williams, D. J., Limburg, W. W., Pearson, J. M., Goedde, A. O., and Yanus, J. F. (1975)
J. Chem. Phys. 62,1501.
Yamamoto, M., Ohoka, M., Kitagawa, K., Nishimoto, S., and Nishijima, Y. (1973) Chem.
Lett. 7 , 745.
Yoshimoto, S., Akana, Y., Kimura, A., Hirata, H., Kusabayashi, S., and Mikawa, H. (1969)
J. Chem. SOC.Chem. Commun. 11,987.
Yoshimura, M., Shirota, Y., and Mikawa, H. (1973) J. Polym. Sci. Polym. Lett. Ed. 11,
457.
Yoshimura, M., Mikawa, H., and Shirota, Y. (1975) Macromolecules 8, 713.

Vous aimerez peut-être aussi