You are on page 1of 24

14 The Influence of Personality on Artistic

and Scientific Creativity

GREGORY J. FEIST

Imagine for a moment, the following conversation between a 4-year-old and her mother:
"Mommy, where did I come from?"
"From your father and me."
"Where did you and Daddy come from?"
At this point, the mother heaves a sigh of relief because no more detailed explanation of
the process is required, and answers, "From your grandparents." The conversation goes on
to a few more generations before the little girl becomes more interested in her toys and
leaves this line of questioning.
Wondering about the nature of creativity is in essence one form of the "Where did I come
from" question, namely "Where did it come from?" The above conversation is an important
one, for two reasons. First, it reminds us that humans are inherently curious about how things
come into being and, second, our answers tend to be reductive. We all wonder were we come
from, and so it is only natural that when we see something we have never seen before we won-
der, "Where did that came from?" The most obvious and superficial answer is, "from its cre-
ator." For some - like the 4-year-old girl - this answer suffices. For most people, however, it
does not. We want to know more specifically about the processes and conditions that allow
one person to create what others never imagined. How one person differs from others is
exactly what is meant by individual differences, and with this emphasis on individual differ-
ences we move close to a personality explanation, the topic of this chapter. Indeed, one pur-
pose of this chapter is to demonstrate that personality is an important answer to the ques-
tions, "Where did that come from and why was she or he able to create it?"
Before doing so, however, let me say a few words about how personality and creativity are
defined. When I tell people I am a personality psychologist, they often wonder what a per-
sonality psychologist is. As I frequently tell my students, there is something specific to what
psychologists mean by the word personality. Psychologists most often define it in terms of
individual differences and behavioral consistency, with the latter coming in two forms: situa-
tional and temporal. Situational consistency focuses on whether people behave consistently
in different situations, whereas temporal consistency focuses on whether they behave con-
sistently over time. Let's take the trait of friendliness, for example. We would only label a per-
son as "friendly" if we observe him or her behaving in a friendly manner over time and in many
different situations, and in situations in which other people did not behave as friendly.
Moreover, if I tell people that my primary interest is creativity, they sometimes argue, as
many academic psychologists have, that creativity by definition is mysterious and beyond the
pale of empirical scrutiny. That may be true concerning the process of creativity, but it fails
to distinguish two other important and observable aspects to creativity, namely the person
and the product. The inner workings of the creative mind may forever be outside of direct
observation, but the behavioral dispositions of the person creating are not. The question,

273

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
274 G. J. FEIST
however, of what is creative behavior and creative work is still unanswered. Psychologists
and philosophers who study the creative process, person, and product are in consensus
about what is "creative": novel and adaptive solutions to problems (Amabile, 1996; Feist,
1993; MacKinnon, 1970; Rothenberg& Hausman, 1976; Simonton, 1988; Sternberg, 1988).
The adaptive criterion is necessary to distinguish truly creative thinking from merely differ-
ent and/or pathological thinking.
Now we can return to the question of how some people are consistently able to come up
with novel and adaptive solutions when most others are not. The most intriguing issue raised
with regard to personality and creativity is the potential causal link between the two
domains. The more general form of the question of whether traits cause behavior, attitudes,
or talents (see McCrae & Costa, 1995) can be framed more specifically: Does personality
cause or influence creative achievement? The primary purpose of this chapter is to demon-
strate that personality has an influence on creative achievement in art and science, by
reviewing the literature on personality and creativity and organizing it around two criteria of
causality: covariance and temporal precedence.
Much contention and acrimony accompanying a term such as cause can be avoided if we
clarify its meaning. Rosenthal and Rosnow (1991) have argued that three criteria must be
met before one can consider a causal relationship between two variables: (a) covariation, (b)
temporal precedence, and (c) ruling out extraneous explanations. Simply put, covariation
concerns the degree to which X covaries with Y If X and Y do not covary, then they cannot
possibly be causally related. Temporal precedence is the idea that X must precede Y in time
if it is to have a causal influence on it. Finally, if we are to conclude that X has a causal influ-
ence on Y we must be able to rule out alternative explanations - X and only X must cause Y
By defining cause in these operational terms, we can address the extent to which personal-
ity "causes" creative achievement. In this sense, "cause" and "influence" can be used inter-
changeably.
Indeed, the first two criteria (covariation and temporal precedence) map nicely onto the
two main components of personality, individual differences and temporal consistency. The
covariation criterion can be examined by studying how individual differences relate to artis-
tic and scientific creative ability. Assuming, however, that the connection between creative
personality and achievement can be demonstrated, is it fair to talk about the "creative per-
sonality" monolithically? The individual difference traits that distinguish creative artists may
not be the same as those that distinguish creative scientists. Therefore, in the individual dif-
ference section of this chapter I will review the literatures for artistic and scientific creativ-
ity separately and then discuss whether the distinguishing traits of the artists generalize to
scientists.
In similar fashion, the temporal consistency of creative personality can shed light on the
temporal precedence criterion. Do the distinguishing traits of creative people measured at
an earlier time in life continue to distinguish them from their peers later in life? If traits do
not distinguish young creative people from their less creative peers, but do so later in life,
then these traits clearly cannot precede creativity. The temporal consistency section, there-
fore, will review longitudinal studies that have examined the consistency of creative person-
ality. Yet a second form of consistency can be explored - that of creative achievement. Are
young gifted people likely to maintain and realize their creative potential by becoming true
creative geniuses in adulthood? In this section I will therefore review longitudinal studies
that have examined whether early creative potential is actualized, with emphasis being
placed on the Terman studies of genius.
In a third section of the chapter I will integrate the first two sections by reviewing theo-
ries that provide underlying mechanisms connecting personality to creative achievement. In

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 275
other words, if there is a connection between personality and creativity, what plausible bio-
logical and/or psychological mechanisms could be responsible for their association? The
recent theories of Eysenck, Russ, Busse and Mansfield, as well as the five-factor model will
be highlighted. In the final section of the chapter I will summarize what we know and do not
yet know after 45 years of empirical research and make suggestions for further exploration.

INDIVIDUAL DIFFERENCES IN CREATIVE PERSONALITY:


EVIDENCE OF COVARIATION

Personality and Artistic Creativity


Before delving into the literature on personality differences in artistic creativity, I need to
say a few words about how artist was defined and to whom artists were compared. For oper-
ational purposes, I defined artist most broadly to include not only visual artists (painters,
sculptors, cinematographers, photographers, architects), but also literary (writers, poets)
and performing artists (musicians, singers, dancers, actors). Students as young as high school
were included if they showed artistic interest and/or promise. Furthermore, in order to
demonstrate that personality meaningfully covaries with artistic creativity, I included stud-
ies in the review only if they compared the personality characteristics of artists versus
nonartists.

NONSOCIAL TRAITS: OPENNESS TO EXPERIENCE, FANTASY, AND IMAGINATION.


Although such a finding may appear self-evident, it is important to document the empirical
support (or lack thereof) for the idea that artists are more open to experience and open to
fantasy and imagination than are nonartists (see Table 14.1). For example, Domino (1974)
studied a group of cinematographers and found that they were quite willing and interested
in seeking out new experiences. More recently, Pufal-Struzik (1992) examined personality
differences between 177 creative artists (painters, poets, writers, and film directors) and
their less creative peers. Using Cattells Sixteen Personality Factor Questionnaire (16 PF),
she found that creative artists were more aesthetically oriented, imaginative, and intuitive.

NONSOCIAL TRAITS: IMPULSIVITY AND LACK OF CONSCIENTIOUSNESS. Closely


related to their rebellious nonconforming dispositions (see later), artists appear to be rather
impulsive and rate low on conscientiousness (see Table 14.1). One illustrative study in this
regard was conducted by Dudek, Berneche Berube, and Royer (1991). Using self-reports on
the Adjective Check List, they examined the personality characteristics of art students and
nonart controls and found that art students were significantly lower on the self-control
(impulsivity) and need for order (conscientiousness) scales. Another example was a study
conducted by Walker, Koestner, and Hum (1995), who investigated the personalities of
highly eminent artists and compared them with eminent controls (political, judicial, and mil-
itary leaders). Personality was assessed by raters reading the last 50 pages of autobiographies
(to focus on adulthood) and rating the subject using the California Q-set (CQS). The CQS
ratings revealed that the artists were significantly more impulsive and less conscientious than
their noncreative peers.

NONSOCIAL TRAITS: ANXIETY, AFFECTIVE ILLNESS, AND EMOTIONAL SENSITIV-


ITY. Another common stereotype of the artist is that of an emotionally labile, manic,
expressive, and sensitive person (see Feist, 1991; Rossman & Horn, 1972; Rothenberg,
1990; Runco & Bahleda, 1986). The real question, however, concerns whether empirical

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Table 14.1. Consistent Personality Findings from the Literature Comparing Artists
and Nonartists

Trait Category Trait Citation

Nonsocial Openness to experience Alter (1989)


Fantasy-oriented Bachtoldfc Werner (1973)
Imagination Barron(1972)
Barton & Cattell (1972)
Cross etal. (1967)
Csikszentmihalyi & Getzels (1973)
Domino (1974)
Eiduson(1958)
Feist (1989)
Getzels & Csikszentmihalyi (1976)
Hall & MacKinnon (1969)
Holland & Baird (1968)
Kemp (1981)
MacKinnon (1962)
Martindale (1975)
Pufal-Struzik(1992)
Rossman & Horn (1972)
Schaefer (1969, 1973)
Shelton & Harris (1979)
Walker etal. (1995)
Zeldow(1973)
Impulsivity Bachtoldfc Werner (1973)
Lack of conscientiousness Bakker(1991)
Barron (1972)
Barton & Cattell (1972)
Cross etal. (1967)
Drevdahlfc Cattell (1958)
Dudeketal. (1991)
Getzels & Csikszentmihalyi (1976)
G6tz&G6tz(1979)
Hall & MacKinnon (1969)
Hammond & Edelmann (1991)
Helson (1977)
Mohan & Tiwana (1987)
Pufal-Struzik(1992)
Schaefer (1969, 1973)
Walker etal. (1995)
Zeldow(1973)
Anxiety Andreasen & Glick (1988)
Affective illness Bakker(1991)
Emotional sensitivity Barron (1972)
Cross etal. (1967)
Csikszentmihalyi & Getzels (1973)
Dudek (1968)
Drevdahl& Cattell (1958)
Eiduson (1958)
Getzels & Csikszentmihalyi (1976)
G6tz&G6tz(1979)
Hall & MacKinnon (1969)
Hammer (1966)
Hammond & Edelmann (1991)
Helson (1977)
Jamison (1993)
Kemp (1981)
Ludwig(1995)
Marchant-Haycox & Wilson (1992)

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Table 14.1 (cont.)

Trait Category Trait Citation

Martindale (1975)
Mohan & Tiwana (1986)
Richards (1994)
Richards & Kinney (1990)
Schaefer (1969,1973)
Shelton & Harris (1979)
Walker etal. (1995)
Wills (1983)
Wilson (1984)
Drive Alter (1989)
Ambition Bakker(1988,1991)
Cross etal. (1967)
Csikszentmihalyi & Getzels (1973)
Domino (1974)
Drevdahl & Cattell (1958)
Dudeketal. (1991)
Eiduson (1958)
Getzels & Csikszentmihalyi (1976)
Hammer (1966)
Helson (1977)
Kemp (1981)
Marchant-Haycox & Wilson (1992)
Schaefer (1969, 1973)
Wilson (1984)
Social Norm doubting Amos (1978)
Nonconformity Bachtoldfc Werner (1973)
Independence Barron (1972)
Barton & Cattell (1972)
Cross etal. (1967)
Csikszentmihalyi & Getzels (1973)
Domino (1974)
Drevdahl & Cattell (1958)
Dudek etal. (1991)
Getzels & Csikszentmihalyi (1976)
Hall & MacKinnon (1969)
Helson (1977)
Holland & Baird (1968)
Kemp (1981)
MacKinnon (1962)
Pufal-Struzik(1992)
Rossman & Horn (1972)
Schaefer (1969,1973)
Shelton & Harris (1979)
Zeldow(1973)
Hostility Barton & Cattell (1972)
Aloofness Cross etal. (1967)
Unfriendliness Drevdahl & Cattell (1958)
Lack of warmth Dudek etal. (1991)
Eysenck(1995)
Getzels & Csikszentmihalyi (1976)
G6tz&Gotz(1979)
Hall & MacKinnon (1969)
Hammond & Edelmann (1991)
Marchant-Haycox & Wilson (1992)
Mohan & Tiwana (1987)
Schaefer (1969,1973)
Wilson (1984)

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
278 G. J. FEIST
research supports or fail to supports this stereotype. In this case, it does. Research suggests
that artists are indeed more emotional and sensitive than nonartists (see Table 14.1). For
instance, Marchant-Haycox and Wilson (1992) administered the Eysenck Personality Pro-
file to 162 performing artists (actors, dancers, musicians, and singers) and found that they
scored significantly higher than control subjects on anxiety, guilt, and hypochondriasis. Sim-
ilarly, Hammond and Edelmann (1991), using the Eysenck Personality Questionnaire,
found that professional actors scored significantly higher on the neuroticism scale than did
nonactor comparison subjects.
A related and rather consistent finding has been the association between artistic creativ-
ity and the disposition toward affective illness. In the most impressive study of its kind, Lud-
wig (1995) examined the relative rates of mental and affective illness in 1,005 eminent peo-
ple in 18 professions. Ludwigs main finding was that all forms of psychopathology (alcohol
and drug abuse, psychosis, anxiety disorders, somatic problems, and suicide, among others)
were more common in the artistic professions than in all other professions. Other research,
however, suggests that bipolar affective disorder is more closely linked with artistic creativ-
ity than is unipolar affective disorder (Andreasen & Glick, 1988; Feist, in press; Jamison,
1993; Richards, 1994; Richards & Kinney, 1990; Russ, 1993). Andreasen and Glick s (1988)
work on a sample of 30 gifted writers and 30 control subjects is illustrative. They found that
writers were more likely to suffer from affective disorder than were control subjects (80%
to 30%) and, more specifically, were more likely to suffer from bipolar disorder (43% to
10%). However, the two groups did not differ on their rates of unipolar depression.
Not all studies have found high levels of anxiety among creative artists (Barton & Cattell,
1972; Buttsworth & Smith, 1994; Feist, 1989; Walker et al., 1995). Buttsworth and Smith
(1994) administered the 16 PF to 255 undergraduate music students and compared them
with 296 undergraduate psychology students. Results revealed that the music students had
more stability (Factor C) and less Anxiety (Factor II) than the psychology students. Simi-
larly, Walker et al. (1995) found that eminent artists were more depressed but not more anx-
ious than their noncreative eminent peers.

NONSOCIAL TRAITS: DRIVE AND AMBITION. One last nonsocial personality character-
istic that tends to distinguish artists from nonartists is drive and ambition (see Table 14.1).
For example, Dudek et al. (1991) reported significantly higher levels of need for achieve-
ment in a sample of professional artists compared with almost 400 nonartist adults. Similarly,
Bakker (1988, 1991) found that adolescent dancers were more achievement oriented and
driven than were an adolescent comparison group.

SOCIAL TRAITS: NORM DOUBTING, NONCONFORMITY, AND INDEPENDENCE. There


are also a set of interpersonal and socially oriented personality traits that are relevant to artis-
tic creativity, one of which is rebellion, or nonconformity. Artists, perhaps more than almost
any other members of society, tend to question and rebel against established norms. Some
may even argue that questioning, challenging, and pushing the limits of what is acceptable
may be the defining traits of being an artist in modern society. The empirical literature on
personality and artistic creativity supports the nonconforming, rebellious nature of artists
(see Table 14.1). For instance, Hall and MacKinnon (1969) found that the most creative
architects scored low on the "communality," "good impression," and "achievement via con-
formance" scales of the California Psychological Inventory (CPI) and scored low on the
"affiliation" scale of the Adjective Check List, but high on its autonomy scale. This pattern
of results depicts personalities that are conflicted, impulsive, nonconformist, rule doubting,
skeptical, independent, and not concerned with obligations or duties. Similarly, research
using the 16 PF suggests that artists are low on Conformity (Factor G), and high on Radi-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 279
calism (Factor Q2) and Self-Sufficiency (Factor Q2). For instance, Barton and Cattell (1972)
and Bachtold and Werner (1973) reported the preceding pattern of results on female sam-
ples of artists. Only one study has reported a negative relationship between being an artist
and radicalism (Buttsworth & Smith, 1994), and it was a study of performance-oriented
music students.

SOCIAL TRAITS: HOSTILITY, ALOOFNESS, UNFRIENDLINESS, AND LACK OF


WARMTH. Research has also pointed toward a cluster of asocial and even antisocial per-
sonality dispositions associated with artistic creativity (see Table 14.1). In one of the earliest
studies of its kind, Drevdahl and Cattell (1958) examined the relationship between artistic
creativity and personality in three samples of artists (writers, visual artists, and science fic-
tion writers). All three groups scored much lower than norms on the 16 PF Factor A
(Warmth). Similarly, Getzels and Csikszentmihalyi (1976) investigated a sample of success-
ful art students and found very low levels of Warmth (Factor A) on the 16 PF. Furthermore,
Eysencks notion of psychoticism, which consists of traits such as aggression, aloofness, anti-
social and egocentric behavior, and tough-mindedness, tends to be higher in artists than
nonartists (Gotz & Gotz, 1979; Hammond & Edelmann, 1991; Mohan & Tiwana, 1987). The
only null result on psychoticism was reported by Wills (1983); the only null result on warmth
was reported by MacKinnon (1962); and the only positive relationship on agreeableness was
reported by Walker et al. (1995).

SOCIAL TRAITS: INTROVERSION. One of the more consistent findings from the person-
ality literature of artists is that they tend to be rather introverted (see Table 14.1). Indeed,
Storr (1988) has argued that the ability to be alone and away from others is a necessary pre-
requisite for creative activity. Only those who make time to be by themselves can spend the
necessary amount of time thinking and creating.
A few studies, however, have reported high levels of extraversion among creative artists,
but these have been with performing artists such as actors or opera singers (Hammond &
Edelmann 1991; Wilson, 1984). For instance, using the Eysenck Personality Questionnaire
and a shyness and sociability scale, Hammond and Edelmann (1991) reported elevated
Extraversion and Sociability scores and depressed Shyness scores in a sample of 51 profes-
sional actors when compared with a sample of 52 nonactors.
In sum, the personality of the creative artist suggests a person who is imaginative, open to
new ideas, drives, neurotic, affectively labile, but for the most part asocial and at times even
antisocial.

Personality and Scientific Creativity


One could argue that science has more latitude of creative expression than art. That is, sci-
entific investigations can range from the very routine, rote, and prescribed to the revolu-
tionary and highly creative breakthrough. In fact, as Kuhn (1970) argued, much of the time
science is relatively mundane and "normal," whereas only rarely does some individual pro-
duce truly "revolutionary science." Granted, some art can be very derivative and merely
technical, yet anyone who makes a living at art must be rather creative. Scientists, on the
other hand, can make a living being little more than technicians. Therefore, the appropriate
comparison group for the creative scientist is the less creative scientist rather than the non-
scientist. Furthermore, I classified a sample in the "science" category if it consisted of either
professionals or students in natural science, biological science, social science, engineering,
invention, or math.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
280 G. J. FEIST
NONSOCIAL TRAITS: OPENNESS TO EXPERIENCE AND FLEXIBILITY OF THOUGHT.
A consistent finding in the personality and creativity in science literature has been that cre-
ative and eminent scientists tend to be more open to experience and more flexible in thought
than are less creative and eminent scientists (see Table 14.2). Many of these findings stem
from data on the flexibility scale (Fe) of the CPI (Feist & Barron, 1996; Garwood, 1964;
Gough, 1961; Helson, 1971; Helson & Crutchfield, 1970; Parloff & Datta, 1965). The Fe
scale taps into flexibility and adaptability of thought and behavior, as well as the preference
for change and novelty (Gough, 1987). The few studies that have reported either no effect
or a negative effect of flexibility in scientific creativity have been with student samples
(Davids, 1968; Smithers & Batcock, 1970).

NONSOCIAL TRAITS: DRIVE, AMBITION, AND ACHIEVEMENT. The most eminent and
creative scientists also tend to be more driven, ambitious, and achievement oriented than
their less eminent peers (see Table 14.2). Busse and Mansfield (1984), for instance, studied
the personality characteristics of 196 biologists, 201 chemists, and 171 physicists, and com-
mitment to work (i.e., "need to concentrate intensively over long periods of time on one s
work") was the strongest predictor of productivity (i.e., publication quantity), even when
holding age and professional age constant. Of course, drive and ambition are predictive of
success in other fields as well, but it is nevertheless important to demonstrate its effect in
science as well. Helmreich, Spence, Beane, Lucker, and Matthews (1980) studied a group
of 196 academic psychologists and found that different components of achievement and
drive had different relationships with objective measures of attainment (i.e., publications
and citations). With a self-report measure, they assessed three different aspects of achieve-
ment: mastery (preferring challenging and difficult tasks), work (enjoying working hard),
and competitiveness (liking interpersonal competition and bettering others). According to
Amabiles (1996) well-known typology, the first two measures could be classified as "intrin-
sic motives" and the last measure could be an "extrinsic motive." Helmreich and his col-
leagues found that mastery and work were positively related to both publication and citation
totals, whereas competitiveness was positively related to publications but negatively related
to citations. Being intrinsically motivated (mastery and work) appears to increase one s pro-
ductivity and positive evaluation by peers (citations), whereas wanting to be superior to
peers leads to increased productivity, and yet a lower positive evaluation by peers. The infer-
ence here is that being driven by the need for superiority may backfire in terms of having a
negative impact on the field. Indeed, in a further analysis their 1980 data set of the male psy-
chologists, Helmreich and colleagues (Helmreich, Spence, & Pred, 1988) factor-analyzed
the Jenkins Activity Survey and extracted an Achievement Striving factor and an Impa-
tience/Irritability factor. Achievement Striving was positively related to both citation and
publication counts, whereas Impatience/Irritability was related to neither publications nor
citations.

SOCIAL TRAITS: DOMINANCE, ARROGANCE, HOSTILITY, AND SELF-CONFIDENCE.


In the highly competitive world of science, especially big science, where the most produc-
tive and influential continue to be rewarded with more and more of the resources, success
is more likely for those who thrive in competitive environments, that is, for the dominant,
arrogant, hostile, and self-confident (see Table 14.2). For example, Van Zelst and Kerr
(1954) collected personality self-descriptions on 514 technical and scientific personnel from
a research foundation and a university. Holding age constant, they reported significant par-
tial correlations between productivity and describing oneself as "argumentative,"
"assertive," and "self-confident." In one of the few studies to examine female scientists,
Bachtold and Werner (1972) administered Cattells 16 PF to 146 women scientists and

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 281

Table 14.2. Consistent Personality Findings from the Literature Comparing Creative and
Less Creative Scientists

Trait Category Trait Citation

Nonsocial Openness to experience Feist & Barron, 1996


Flexibility of thought Garwood, 1964
Gough, 1961
Helson, 1971
Helson & Crutchfield, 1970
Parloff&Datta, 1965
Parloff, Datta, Kleman, & Handlon, 1968
Roco, 1993
Rossman & Horn, 1972
Schaefer, 1969
Shapiro, 1968
Van Zelst & Kerr, 1954
Wispe, 1963
Drive Albert & Runco, 1987
Ambition Bloom, 1956
Achievement Busse & Mansfield, 1984
Chambers, 1964
Davids, 1968
Erickson et al., 1970
Feist, 1993
Gantz, Erickson, & Stephenson, 1972
Gough, 1961
Helmreich et al., 1980
Helmreich et al., 1988
Holland, 1960
Ikpaahindi, 1987
Lacey & Erickson, 1974
Rushton et al., 1983
Schaefer, 1969
Shapiro, 1968
Simon, 1974
Van Zelst & Kerr, 1954
Wispe, 1963
Social Dominance Bachtold & Werner, 1972
Arrogance Chambers, 1964
Hostility Davids, 1968
Self-confidence Erickson et al., 1970
Feist, 1993
Gantz et al., 1972
Garwood, 1964
Gough, 1961
Ham & Shaughnessy, 1992
Helmreich et al., 1988
Helson & Crutchfield, 1970
Lacey & Erickson, 1974
McDermid, 1965
Parloff&Datta, 1965
Parloff, et al., 1968
Rossman & Horn, 1972
Rushton et al., 1983
Schaefer, 1969
Shapiro, 1968
Van Zelst & Kerr, 1954
Wispe, 1963

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
282 G. J. F E I S T

Table 14.2 (cont)

Trait Category Trait Citation

Autonomy Albert & Runco, 1987


Introversion Bachtold & Werner, 1972
Independence Bloom, 1956
Busse & Mansfield, 1984
Chambers, 1964
Davids, 1968
Erickson et al., 1970
Garwood, 1964
Helson, 1971
Helson & Crutchfield, 1970
Holland, 1960
Lacey & Erickson, 1974
Parloff&Datta, 1965
Roco, 1993
Roe, 1952
Rossman & Horn, 1972
Rushton et al., 1987
Schaefer, 1969
Smithers & Batcock, 1970
Terman, 1955
Van Zelst & Kerr, 1954

found that they were significantly different from women in general on 9 of the 16 scales,
including Dominance (Factor E) and Self-confidence (Factor O). Similarly, Feist (1993)
recently reported a structural equation model of scientific eminence in which the path
between observer-rated hostility and eminence was direct and the path between arrogant
working style and eminence was indirect but significant.

SOCIAL TRAITS: AUTONOMY, INTROVERSION, AND INDEPENDENCE. The scientific


elite also tend to be more aloof, asocial, and introverted than their less creative peers (see
Table 14.2). In a classic study concerning the creative person in science, Roe (1952, 1953)
found that creative scientists were more achievement oriented and less affiliative than were
less creative scientists. In another seminal study of the scientific personality, Eiduson (1962)
found that scientists were independent, curious, sensitive, intelligent, emotionally invested
in intellectual work, and relatively happy. Similarly, Chambers (1964) reported that creative
psychologists and chemists were markedly more dominant, ambitious, and self-sufficient,
and had more initiative as compared with less creative peers. Helson (1971) compared cre-
ative female mathematicians with less creative female mathematicians, matched on IQ.
Observers blindly rated the former as having more "unconventional thought processes," as
being more "rebellious and non-conforming," and as being less likely to judge "self and oth-
ers in conventional terms." More recently, Rushton, Murray, and Paunonen (1987) con-
ducted factor analyses of the personality traits most strongly loading on the Research factor
(in contrast to a Teaching factor) in two separate samples of academic psychologists. Among
other results, they found that "independence" tended to load on the Research factor,
whereas "extraversion" tended to load on the Teaching factor.
To summarize the distinguishing traits of creative scientists: They are generally more
open and flexible, driven and ambitious, and although they tend to be relatively asocial,

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 283
when they do interact with others, they tend to be somewhat prone to arrogance, self-
confidence, and hostility.

Generalizability of Personality Traits of Creative People


Historically most reviews of the creativity and personality literature have either failed to dis-
tinguish between artistic and scientific creativity or have not explicated the unique charac-
teristics of each domain (Barron & Harrington, 1981; Delias & Gaier, 1970; Mumford &
Gustafson, 1988; Stein, 1968). Given the accumulation of findings on personality and cre-
ativity, however, we are now in a position to be more systematic and discriminating with the
trends and patterns that have developed. One way of doing so is by making explicit the sim-
ilarities and differences between the different creative domains, in particular, art and sci-
ence.

Personality Characteristics Unique to Creative Artists and Scientists


Compared with creative scientists, artists appear to be more anxious, emotionally labile, and
impulsive. More generally, therefore, the artistically creative person appears to have a dis-
position toward intense affective experience (Andreasen & Glick, 1988; Bamber, Bill, Boyd,
& Corbett, 1983; Csikszentmihalyi & Getzels, 1973; Gardner, 1973; Getzels & Csikszentmi-
halyi, 1976; Jamison, 1993; Ludwig, 1995; Richards, 1994; Russ, 1993; Simonton, 1988). To
quote Russ (1993), "One of the main differences between artistic and scientific creativity
may be the importance of getting more deeply into affect states and thematic material in
artistic creativity" (p. 67). To the extent that art is more often an introspective journey and
science more of an externally focused one (see Gardner, 1973), it is not surprising that artists
would be more sensitive to and expressive of internal emotional states than are scientists.
This is not to say that the creative process in art is exclusively emotional and in science exclu-
sively nonemotional. Research suggests that that is not the case (Feist, 1991). The discovery
stages of scientific creativity are often very intuitive and emotional, just as the elaboration
stages of artistic creativity can be very technical and tedious. Yet dispositionally, artists and
scientists generally do tend to differ on the degree to which they are sensitive to their own
and other people s emotional states.
The second core set of unique characteristics of the artistic personality can be classified
as low socialization and low conscientiousness. Although it is true that low socialization and
nonconformity are traits of both creative artists and scientists (Barron, 1963,1972; Cattell &
Drevdahl, 1955; Csikszentmihalyi & Getzels, 1973; Hall & MacKinnon, 1969; Helson, 1971;
Kemp, 1981; Ochse, 1990), the form that nonconformity takes may be different in the two
professions. Artists and not scientists, for instance, tend to be much lower than the norm on
the "socialization," "communality," "tolerance," and "responsibility" scales of the CPI (e.g.,
Barron, 1972; Domino, 1974; Zeldow, 1973) and the Qx (Radical) scale of the 16 PF (Csik-
szentmihalyi & Getzels, 1973; Drevdahl & Cattell, 1958; Kemp, 1981). The low socialization
and responsibility scores are indicative of people who very much question, doubt, and strug-
gle with social norms. Perhaps artists are more actively nonconformist or asocialized than are
scientists, who may be less overt in their nonconformity. Consistent with the idea that sci-
entists are less actively nonconforming than are artists, scientists in general tend to be more
conscientious and orderly than are nonscientists (Kline & Lapham, 1992; Rossman & Horn,
1972; Schaefer, 1969; Wilson & Jackson, 1994). When one considers that traits such as being
"organized," "planful," "not careless," and "not slipshod" make up the conscientiousness

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
284 G. J. FEIST
dimension (John, 1990), it is not surprising that scientists would score higher on the dimen-
sion than could artists.

Personality Commonalities Between Creative Artists and Scientists


If emotional lability, impulsivity, nonconformity, and rebelliousness tend to distinguish
artists from scientists, then what personality characteristics do creative artists and scientists
tend to have in common?
The evidence suggests that the creative person (artist or scientist) in general is distin-
guished by relatively high levels of asocial characteristics, namely introversion, indepen-
dence, hostility, and arrogance (Bachtold & Werner, 1972; Csikszentmihalyi & Getzels,
1973; Diidek et al., 1991; Feist, 1993, 1994; Garwood, 1964; Guastello & Shissler, 1994;
Kline & Lapham, 1992; Ochse, 1990; Rushton et al., 1987; Storr, 1988; Zeldow, 1973). Many
have argued that isolation, withdrawal, and independence are necessary conditions of cre-
ative achievement (Barron, 1963, 1972; Csikszentmihalyi & Getzels, 1973; Ochse, 1990;
Storr, 1988). For instance, Storr (1988) argued that present-day Western society overem-
phasizes interpersonal relationships as the source of happiness and well-being and under-
emphasizes the solitude of creative achievement. In addition, Feist (1993,1994) has demon-
strated the prevalence of hostility among highly creative scientists. In his model of scientific
eminence, observer ratings of hostility and arrogance each had direct and indirect effects on
eminence (Feist, 1993). Furthermore, the personality constellations of scientists who think
complexly about research were quite distinct from those who think complexly about teach-
ing (Feist, 1994). Specifically, the former were seen by others as more hostile and exploita-
tive, whereas the latter were viewed as more gregarious and warm.
A second cluster of distinguishing traits revolve around the need for power and for diver-
sity of experience: drive, ambition, self-confidence, openness to experience, flexibility of
thought, and active imagination. In order to achieve and to go against the norms, one must
have a rather high energy level and be driven (see Amabile, 1996; Barron, 1963, 1972; Shel-
don, 1995; Sternberg & Lubart, 1995). Belief in what one is doing and the ambition to do it
originally are probably related to the high degree of self-confidence often seen in creative
individuals. It is then only a short step from drive and self-confidence to arrogance and hos-
tility. If one is intrinsically driven by a task and has a need to be alone, unbothered by oth-
ers, as is often true of creative people, then social approval and social niceties are not likely
to be high on one s list of priorities. Hostility and arrogance may, therefore, be a result of
complete dedication and devotion to work, and anything that detracts from that work would
be the object of scorn and hostility.

TEMPORAL CONSISTENCY IN CREATIVE PERSONALITY: EVIDENCE OF


TEMPORAL PRECEDENCE

Consistency of Creative Personality


The only methodology that provides a possible answer to the temporal precedence (which
comes first) question is longitudinal research. Only longitudinal studies can speak to
whether the distinguishing traits of creative people measured at an earlier time in life con-
tinue to distinguish them from their peers later in life. Showing that traits such as indepen-
dence, introversion, openness, hostility, and dominance exist in high levels early in life may
not mean that they precede creativity, but such a pattern of results would be consistent with
precedence of personality. On the other hand, if these are not the traits that distinguish
young creative people from their less creative peers, but do so later, then they clearly can-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 285
not precede creativity. Are there any longitudinal studies that find a set of personality traits
that distinguish young creative people from their less creative peers but do not later in life?
No, the literature overwhelmingly points toward the consistency of creative personality
(Camp, 1994; Dudek et al., 1991; Dudek & Hall, 1991; Feist, 1995; Getzels & Csikszentmi-
halyi, 1976; Helson, 1987; Helson, Roberts, & Agronick, 1995; Schaefer, 1973; Stohs, 1990;
Terman, 1954). For example, Schaefer (1973) conducted a five-year follow-up investigation
of creative young adults who were originally tested in adolescence. The adolescent sample
consisted of 100 participants in each of the following four criterion groups: creative art/
writing males, creative science males, creative art females, and creative writing females.
There were also 100 participants in four matched control groups. Roughly half of each sam-
ple participated in a replication five years later. Many of the same scales (i.e., autonomy, self-
control, and nurturance) that distinguished creative adolescents from their peers continued
to distinguish the two groups in early adulthood. In other research on the consistency of cre-
ative personality, Dudek and Hall (1991) studied three groups of architects and concluded
that "it is evident that Group III [the less creative architects] retained its social conformity
and Group I [the creative architects] its spontaneity and independence over the 25 years"
(p. 218). Finally, Helson, Roberts, and Agronick (1995) found that creative women at age 52
were consistently rated by observers at age 21 and age 43 as being aesthetically oriented,
interesting, driven, rebellious, independent, and not conventional, conservative, or submis-
sive.

Consistency of Creative Achievement


We have now seen that the distinguishing personality traits of creative people tend to be
rather stable from adolescence or early adulthood on. It is also possible that possessing cer-
tain personality characteristics may play a vital role in determining which gifted and talented
children go on to actualize their potential and which do not. Indeed, although there is a
mounting body of evidence that intellectual precocity and giftedness foreshadow educa-
tional attainment and career success, they do not appear to systematically foreshadow adult
creative achievement. Let me be more specific. Gifted children (i.e., those with high IQs)
are more likely to be more successful in school, obtain higher degrees, and are more likely
to go into professional and/or business professions than are less gifted students (Benbow &
Minor, 1986; Benbow & Stanley, 1982; Cox, 1926; Holahan & Sears, 1995; Lubinski & Ben-
bow, 1994; Pyryt, 1992; Tomlinson-Keasey & Keasey, 1993; Wise, Steel, & McDonald,
1979). A surprisingly high proportion of gifted students, however, do not continue in careers
in which they exhibited precocious talents or do not make significant creative contributions
to their field as adults (Arnold, 1992; Barron & Harrington, 1981; Cramond, 1994; Farmer,
1988; Gough, 1976; Guilford, 1959; Helson, 1987; Hudson, 1958; MacKinnon, i960; Mar-
land, 1972; Milgram & Hong, 1994; Simonton, 1988; Sternberg, 1988; Subotnik, Duschl, &
Selmon, 1993; Subotnik & Steiner, 1992; Tannenbaum, 1983; Taylor, 1963; Winner & Mar-
tino, 1993). The study by Milgram and Hong (1994) is a good example. They studied a group
of high school students over an 18-year period and found that while grades were related to
academic success, they were not related to any adult accomplishment measures, including a
creative outcome measure. Creative thinking at age 18, however, did predict work accom-
plishment at age 36. As Farmer (1988) also pointed out, only 42% of the male and 22% of
the female extremely precocious students went on to choose science or math graduate pro-
grams (cf. Benbow, 1988; Benbow & Lubinski, 1993). In addition, Subotnik and Steiner
(1992) reported in a group of high school Westinghouse Science winners that only five years
after being so designated, almost 20% of the gifted young males and nearly 40% of the gifted
young female scientists were not pursuing scientific careers. Finally, in one of the few Ion-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
286 G. J. FEIST
gitudinal studies of musical precocity, Winner and Martino (1993) found that early musical
talent often did not translate into adult creative achievement in music.
If giftedness is a poor predictor of creative achievement, then are there any consistent
psychological processes that suggest who will continue on or who will make creative contri-
butions? There is some evidence that gender (females being more likely to drop out) and
motivation (Subotnik, Duschl, & Selmon, 1993) predict creative achievement, but certain
personality traits also play a role (Albert, 1991; Butler-Por, 1993; Helson, 1987; Lindsay,
1978; Tomlinson-Keasey & Keasey, 1993; Tomlinson-Keasey & Little, 1990; Trost, 1993).
For instance, Helson (1987) divided into two groups subjects who in college were nomi-
nated by faculty members to have the highest potential for creative achievement. The two
groups were composed of those who were and were not in creative and successful careers
(writers, dancers, artists, psychotherapists, etc.) at age 43. The former were labeled "suc-
cessful careerists," whereas the latter were labeled "other nominees." Personality data were
collected on both groups at ages 21, 27, and 43, so prospective and contemporary personal-
ity comparisons could be made between the successful careerists and the other nominees.
Helson found that, at age 21, the successful careerists were more dominant, independent,
achievement-oriented, self-accepting, and psychologically minded but were less sociable
than the other nominees. At age 27 they were still more achievement-oriented and psycho-
logically minded and less sociable. Andfinally,at age 43 the successful careerists were more
independent, empathic, status-oriented, self-accepting, responsible, and achievement-
oriented, and still less sociable than other nominees.
Without a doubt the most ambitious and extensive longitudinal study of giftedness has
been the one begun by Lewis Terman in the 1920s. More than 1,500 children with IQs
greater than 135 were followed over the course of their entire lives (in fact, the study is still
being conducted). Extensive follow-ups were conducted on average every 10 years. One rel-
evant finding from the Terman study has been that adolescent personality predicts later edu-
cational attainment (Tomlinson-Keasey & Keasey, 1993; Tomlinson-Keasey & Little, 1990).
Using factor analysis and structural equations models, and taking advantage of the multiple
measures and very large sample size of the study, Tomlinson-Keasey and Little (1990) found
that adolescent sociability and social responsibility (empathy) had direct influences on edu-
cational attainment. Moreover, intellectual determination (i.e., drive or achievement moti-
vation) had a direct impact on maintaining intellectual skill. Tomlinson-Keasey and Keasey
(1993) also reported that for gifted women (i.e., holding IQ constant at a mean of 148),
observer personality ratings made during adolescence were predictive of educational attain-
ment. Adolescent girls who were rated by teachers and parents as being highest on intellec-
tual determination and social responsibility (i.e., empathy conscientiousness) were most
likely to obtain college and postgraduate degrees.
In sum, to the question of whether adolescent talent translates into adult success, the
answer is a rather clear yes. There is little doubt that highly intelligent children are more
likely than their classmates to obtain higher grades in school, obtain higher educational
degrees, and in general go into well-paying professional careers. But are they more likely to
be truly creative? The answer appears to be no. Surprisingly often, the most talented chil-
dren and adolescents go on to lead relatively uncreative lives. Indeed, precocious intellec-
tual talent may be neither necessary nor sufficient for true creative achievement in adult-
hood.
Such a lack of predictive validity of intellectual aptitude tests can be explained by the small
relationship between intelligence and creativity (Barron & Harrington, 1981; Getzels, 1987;
MacKinnon, 1978; Magnusson & Backteman, 1978; Milgram & Hong, 1994; Rossman &
Horn, 1972; Sternberg, 1986; Sternberg & Lubart, 1995; Wallach, 1970; Winner & Martino,
1993). To quickly solve multiple choice problems that have known solutions involves conver-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 287
gent or analytical thinking skills, whereas to creatively solve open-ended problems that have
no known solutions involves divergent or intuitive thinking skills (Guilford, 1950,1959,1987;
Simonton, 1988, 1989; Sternberg, 1986). Creativity is fluency, flexibility, usefulness, and
originality of association, not speed at solving verbal and/or mathematical multiple-choice
problems. It should come as no surprise, therefore, that creative potential and creative abil-
ity are better predictors of later creative achievement than is intellectual ability (Barron &
Harrington, 1981; Milgram & Hong, 1994; Rossman & Horn, 1972; Sternberg & Lubart,
1995)-

INTEGRATIVE THEORY LINKING PERSONALITY AND CREATIVITY

If personality does covary with creative achievement and there is temporal stability of cre-
ative personality, then do we have theoretical models connecting personality and creativity?
Furthermore, are there plausible and testable biological and/or psychological mechanisms
linking the two? Recently, a number of theoretical models have begun to propose plausible
and testable explanations for the connection. Recall that the most consistent correlates of
creativity in both art and science are introversion, drive, ambition, openness, flexibility,
autonomy or introversion, and hostility and arrogance. A few theories have attempted to
connect each of them to creative achievement.
Perhaps the most ambitious and inclusive recent theory of personality and creativity is the
one offered by Eysenck (1993,1995). Eysenck has proposed a causal theory of creativity that
begins with genetic determinants, hippocampal formation (of dopamine and serotonin),
cognitive inhibition, and psychoticism, which in turn leads to trait creativity and ultimately
creative achievement. The most appealing aspect of this model, although speculative in
parts, is that it is testable. What is of particular interest in Eysenck's model are the relation-
ships between genetic and neurochemical processes and trait creativity (i.e., personality),
which is the direct precursor to creative achievement. For instance, a key component impli-
cated in Eysenck's biologically based model is cortical arousal. High arousal is associated
with a narrowing of attention, whereas low arousal is associated with a widening of attention.
Furthermore, Eysenck and other researchers have found that creativity depends on a wide
attentional focus and an expansion of cognitive searching to the point of overinclusion, a
defining characteristic of psychoticism (Eysenck, 1995; Isen, Daubman, & Nowicki, 1987;
Jamison, 1993; Mendelsohn, 1976). Therefore, one would predict that creative thinking
might be related to low cortical arousal. Colin Martindale has established a research pro-
gram that has tested this idea systematically and has consistently found support for it (Mar-
tindale, 1981; Martindale & Armstrong, 1974; Martindale & Greenough, 1973; Martindale
& Hasenfus, 1978; Martindale, Hines, Mitchell, & Covello, 1984). For example, as mea-
sured by stress, high arousal reduces creative solutions to problems (Martindale & Gree-
nough, 1973), and, as measured by EEG (percent time spent in alpha states), low arousal
was related to more creative problem solving (Martindale & Armstrong, 1974). However,
low cortical arousal is evident only during the inspiration stage and not throughout creative
insight or during baseline measures. In fact, creative individuals tend to have higher resting
arousal levels (Martindale & Armstrong, 1974), which is consistent with the high cortical
arousal of introversion and its relation to creativity (Eysenck, 1990, 1995).
More generally, trying to explain the connection between creativity and psychoticism,
Woody and Claridge (1977) wrote: "Both [psychoticism and creativity] may tap a common
factor associated with the willingness to be unconventional or engage in mildly antisocial
behaviour" (p. 247). As mentioned earlier, radical, unconventional, asocial, or even antiso-
cial behaviors are probably more common among artists than scientists, but these traits are
nonetheless elevated in creative scientists relative to norms (Bachtold, 1976; Barton & Cat-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
288 G. J. FEIST
tell, 1972; Getzels & Csikszentmihalyi, 1976; Helson, 1971; Rushton, 1990; Rushton, Mur-
ray, & Paunonen, 1983; Wilson & Jackson, 1994). Whether unconventionality is antecedent
to or consequent of creativity is of course open to question.
However, it is not unbridled psychoticism that is most strongly associated with creativity,
but psychoticism tempered by high ego strength or ego control. Paradoxically, creative peo-
ple appear to be simultaneously very labile and unstable and yet can be rather controlled and
stable (Barron, 1963; Eysenck, 1995; Feist, in press; Fodor, 1995; Richards, Kinney, Lunde,
Benet, & Merzel, 1988; Russ, 1993). As Barron (1963) argued over 30 years ago: "Thus the
creative genius may be at once naive and knowledgeable, being at home equally to primitive
symbolism and to rigorous logic. He is both more primitive and more cultured, more
destructive and more constructive, occasionally crazier and yet adamantly saner, than the
average person" (p. 224).
In addition, Russ (1993) elaborated on the facilitative effect that affective traits play in the
creative process and developed a model that conceptually integrates much of the known
empirical findings concerning the relationship between creativity and affective dispositions.
For instance, she hypothesized that access to affect-laden thoughts (primary process thought
and affective fantasy) and openness to affective states leads to the divergent-thinking abili-
ties of free association, breadth of attention, and fluidity of thought, as well as to the trans-
formation abilities of shifting sets and cognitive flexibility. These paths are essentially the
same as those Eysenck proposed connecting affective states, overinclusive thinking, and cre-
ativity. Furthermore, Russ suggested that taking affective pleasure in challenge and being
intrinsically motivated results in an increased sensitivity to problems and problem finding.
Being sensitive, open, and flexible in thought are in turn important personality dispositions
related to creativity. In sum, there are theoretical and empirical reasons for recognizing the
connection between affective states, affective traits, and creative ability and achievement.
Although coming from the context of scientific creativity, Mansfield and Busse (1981) also
developed an integrated model of personality and creativity (cf. Helmreich et al., 1980).
Their model includes not only paths between personality and creativity, but also develop-
mental antecedents as precursors of personality. Based on empirical findings, they suggested
that particular developmental antecedents precede personality characteristics, which in turn
precede the creative process. The developmental antecedents associated with creative peo-
ple are low emotional intensity of the parent-child relationship, parental fostering of auton-
omy, parental intellectual stimulation, and apprenticeship. These are antecedent to the per-
sonality traits of autonomy, flexibility and openness, need to be original, commitment to
work, need for professional recognition, and, finally, aesthetic sensitivity. Finally, Mansfield
and Busse proposed that these traits facilitate the crucial stages involved in creative achieve-
ment: selection of the problem, extended effort working on the problem, setting constraints,
changing constraints, and, finally, verification and elaboration. One interesting yet difficult
to support assumption of their model is that personality precedes the development of cre-
ativity.
The field of personality psychology has recently witnessed the widespread adoption of the
five-factor model (FFM; Digman, 1990; McCrae & John, 1992), which argues that there are
five fundamental bipolar dimensions to personality: openness, neuroticism, extraversion,
agreeableness, and conscientiousness. Although few researchers have directly examined the
relationship between creativity and the FFM (Dollinger & Clancy, 1993; McCrae, 1987;
Mumford, Costanza, Threlfall, Baughman, & Reiter-Palmon 1993), enough work has accu-
mulated on separate FFM dimensions and creativity that I can summarize the consistent
trends. The strongest relationship exists between creativity and openness, but relationships
also have been reported between each of the other four dimensions and creativity: neuroti-
cism (Andreason & Glick, 1988; Bakker, 1991; Hammond & Edelmann, 1991; Kemp, 1981;

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 289
Marchant-Haycox & Wilson, 1992), lack of conscientiousness (Drevdahl & Cattell, 1958;
Getzels & Csikszentmihalyi, 1976; Kemp, 1981; Shelton & Harris, 1979; Walker et al.,
1995), introversion (Bachtold & Werner, 1973; Busse & Mansfield, 1984; Chambers, 1964;
Cross, Cattell, & Butcher, 1967; Helson, 1971, 1977; Pufal-Struzik, 1992; Roco, 1993; Ross-
man & Horn, 1972; Rushton et al., 1987; Zeldow, 1973), and lack of agreeableness (Barton
& Cattell, 1972; Dudek et al., 1991; Eysenck, 1995; Feist, 1993, 1994; Getzels & Csikszent-
mihalyi, 1976; Hall & MacKinnon, 1969; Helmreich et al., 1988; Helson & Crutchfield,
1970; Lacey & Erickson, 1974; McDermid, 1965). This is not to say that all who have
explored the relationship between the FFM and creativity have found each personality
dimension to relate to creativity (Dollinger & Clancy, 1993; Feist, 1989; McCrae, 1987;
Woody & Claridge, 1977). However, many of these null or negative results were conducted
on general population samples and not on creative artists or scientists. Although not yet
established empirically, it may be that the five factors are more consistently related to artis-
tic and scientific creativity than to everyday creativity. Future research must be conducted,
however, before such a conclusion can be made. At the very least, however, openness to
experience is related to creativity, so how do we account for this association? McCrae (1987)
suggests there were three possible reasons for the link. First, open people may be more fas-
cinated with the open-ended, creative, problem-solving tasks and they may simply score
higher on such tasks. Second, open people may have developed cognitive skills associated
with creative, divergent thinking, namely flexibility and fluidity of thought. And third, open
people may have an interest in seeking sensation and more varied experiences, and this
experiential base may serve as the foundation for flexibility and fluency of thinking. Again,
more research is needed to determine the validity of these speculations.

CONCLUSIONS

Fascination with creative genius has compelled many of greatest minds of Western culture
to put pen to paper, including but not limited to Socrates, Plato, Aristotle, Kant,
Wordsworth, Coleridge, Poe, Galton, Russell, Poincare, Freud, Bergson, Einstein, Maslow,
Rogers, and Skinner (see Ghiselin, 1952; Rothenberg & Hausman, 1976; Vernon, 1970). As
Steinberg and Lubart (1995, 1996) have recently pointed out, few topics are of greater
importance to psychology than creativity. Universities, businesses, the arts, entertainment,
and politics - in other words, all of the major institutions of modern society - are each driven
by their ability to create and solve problems originally and adaptively, that is, creatively.
Therefore, the ultimate success and survival of these institutions depend on their ability to
attract, select, and maintain creative individuals. So what do we know about the creative per-
sonality after 45+ years of systematic empirical work? Table 14.3 summarizes what we know
and what this chapter has presented. Certain personality traits consistently covary with cre-
ativity, yet there are some domain specificities; the temporal stability of creative personality
is consistent with the idea that the distinguishing personality traits may precede creative
achievement; finally, giftedness as measured by IQ tests tends not to be a very valid predic-
tor of adult creative achievement. Some of these conclusions may appear to be deceptively
simple, but they go far in demonstrating that personality as a construct and its study as a dis-
cipline offer a unique and important perspective on creativity and the creative process.

Gaps in the Literature and Work for the Future


There is still a long way to go before consensus can be reached on some of the more diffi-
cult and pressing problems of the creative personality. If we have begun to establish covari-
ation between personality and creativity, we have only begun to address temporal prece-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
290 G. J. FEIST

Table 14.3. Summary of Major Conclusions

Creative people in art and science tend to be open to new experiences, less conventional and less con-
scientious, more self-confident, self-accepting, driven, ambitious, dominant, hostile, and impulsive.
Creative people in art and science do not share the same unique personality profiles: Artists are more
affective, emotionally unstable, as well as less socialized and accepting of group norms, whereas sci-
entists are more conscientious.
The traits that distinguish creative children and adolescents tend to be the ones that distinguish cre-
ative adults. The creative personality tends to be rather stable.
Childhood academic intelligence (giftedness) is a relatively poor predictor of adult creative achieve-
ment.

dence and have all but ignored ruling out extraneous variables as explanations. For instance,
no one has begun systematic investigation of creative potential and ability in young children
and followed them through adolescence and adulthood. Such research has been conducted
on intelligence and giftedness (see, e.g., Subotnik & Arnold, 1994), but not creativity. How
stable is creativity from early childhood to adulthood? Are creativity and intelligence always
distinct or do they diverge only after a certain age? How do the dispositions toward origi-
nality interact with the other psychological processes important to creative achievement,
namely development, cognition, and social influence? Finally, do other psychological
processes account for the correlations between personality and creativity? Only once these
questions are examined systematically and empirically can the theoretical models of the cre-
ative person be evaluated, tested, and modified (Eysenck, 1993, 1995; Feist & Gorman,
1998; Helmreich et al., 1980; Mansfield & Busse, 1981).
Empirical research over the past 45 years makes a rather convincing case that creative
people behave consistently over time and situation and in ways that distinguish them from
others. The creative personality does exist and personality dispositions regularly and pre-
dictably relate to creative achievement in art and science. How the two develop, however,
and their influence on one another for the most part are little understood. Perhaps in the
next 45 years someone with the requisite constellation of personality characteristics and cre-
ative ability will focus his or her energies on the topic and come up with a creative solution.

NOTE
I am grateful to Erika Rosenberg and John Nezlek for their comments on an earlier draft of this chap-
ter. Preparation of this chapter was supported in part by a grant from the Committee on Faculty
Research at the College of William and Mary.

REFERENCES
Albert, R. S. (1991). People, processes, and developmental paths to eminence: A developmental-
interactional model. In R. M. Milgram (Ed.), Counseling gifted and talented children: A guide for
teachers, counselors, and parents (pp. 75-93). Norwood, NJ: Ablex.
Albert, R. S., & Runco, M. (1987). The possible different personality dispositions of scientists and non-
scientists. In D. N. Jackson and J P Rushton (Eds.), Scientific excellence (pp. 67-97). Beverly Hills,
CA: Sage.
Alter, J. B. (1989). Creativity profile of university and conservatory music students. Creativity Research
Journal, 2, 184-195.
Amabile, T (1996). Creativity in context. New York: Westview.
Amos, S. P. (1978). Personality differences between established and less-established male and female
creative artists. Journal of Personality Assessment, 42, 374-377.
Andreasen, N. C , & Glick, L. D. (1988). Bipolar affective disorder and creativity: Implications and clin-
ical management. Comprehensive Psychiatry, 2Q, 207-216.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 291
Arnold, K. D. (1992). Undergraduate aspirations of career outcomes of academically talented women:
A discriminant analysis. Roeper Review, 15, 169-175.
Bachtold, L. M. (1976). Personality characteristics of women of distinction. Psychology of Women
Quarterly, 1, 70-78.
Bachtold, L. M., & Werner, E. E. (1972). Personality characteristics of women scientists. Psychological
Reports, 31, 391-396.
Bachtold, L. M., & Werner, E. E. (1973). Personality characteristics of creative women. Perceptual and
Motor Skills, 36, 311-319.
Bakker, F. C. (1988). Personality differences between young dancers and non-dancers. Personality and
Individual Differences, 9, 121-131.
Bakker, F. C. (1991). Development of personality in dancers: A longitudinal study. Personality and Indi-
vidual Differences, 12, 671-681.
Bamber, J. H., Bill, J. M., Boyd, F. E., & Corbett, W. D. (1983). In two minds: Arts and science differ-
ences at sixth-form level. British Journal of Educational Psychology, 53, 222-233.
Barron, F. (1963). Creativity and psychological health. New York: Van Nostrand.
Barron, F. (1972). Artists in the making. New York: Seminar Press.
Barron, F , & Harrington, D. (1981). Creativity, intelligence, and personality. Annual Review of Psy-
chology, 32, 439-476.
Barton, K., & Cattell, H. (1972). Personality characteristics of female psychology, science and art
majors. Psychological Reports, 31, 807-813.
Benbow, C. P. (1988). Sex differences in mathematical reasoning ability in intellectually talented pread-
olescents: Their nature, effects, and possible causes. Behavioral and Brain Sciences, 11, 169-183.
Benbow, C. P., & Lubinski, D. (1993). Psychological profiles of the mathematically talented. Some sex
differences and evidence supporting their biological basis. In G. R. Bock and K. Ackrill (Eds.), The
origins and development of high ability (pp. 44-66). Chichester: Wiley.
Benbow, C. P., & Minor, L. L. (1986). Mathematically talented students and achievement in the high
school sciences. American Educational Research Journal, 23, 425-436.
Benbow, C. P., & Stanley, J. C. (1982). Consequences in high school and college of sex differences in
mathematical reasoning ability: A longitudinal perspective. American Educational Research Journal,
19> 598-622.
Bloom, B. S. (1956). Report on creativity research at the University of Chicago. In C. W. Taylor (Ed.),
The 1955 University of Utah Research Conference on the Identification of Creative Scientific Talent.
Salt Lake City: University of Utah Press.
Busse, T. V, & Mansfield, R. S. (1984). Selected personality traits and achievement in male scientists.
Journal of Psychology, 116, 117-131.
Butler-Por, N. (1993). Underachieving gifted students. In K. A. Heller, F. J. Monks, & A. H. Passow
(Eds.), International handbook of research and development of giftedness and talent (pp. 649-688).
Oxford: Pergamon.
Buttsworth, L. M., & Smith, G. A. (1994). Personality of Australian performing musicians by gender and
by instrument. Personality and Individual Differences, 5, 595-603.
Camp, G. C. (1994). A longitudinal study of correlates of creativity. Creativity Research Journal, 7,
125-144.
Cattell, R. B., & Drevdahl, J. E. (1955). A comparison of the personality profile (16 PF) of eminent
researchers with that of eminent teachers and administrators, and the general population. British
Journal of Psychology, 46, 248-261.
Chambers, J. A. (1964). Relating personality and biographical factors to scientific creativity. Psycholog-
ical Monographs: General and Applied, 78, 1-20.
Cox, C. (1926). Genetic studies of genius: Vol. 2. The early mental traits of three hundred geniuses. Stan-
ford, CA: Stanford University Press.
Cramond, B. (1994). The Torrance Tests of Creative Thinking: From design through establishment of
predictive validity. In R. F. Subotnik & K. D. Arnold (Eds.), Beyond Terman: Contemporary longitu-
dinal studies of giftedness and talent (pp. 229—254). Norwood, NJ: Ablex.
Cross, P. G., Cattell, R. B., & Butcher, H. J. (1967). The personality pattern of creative artists. British
Journal of Educational Psychology, 37, 292-299.
Csikszentmihalyi, M., & Getzels, J. W. (1973). The personality of young artists: An empirical and theo-
retical exploration. British Journal of Psychology, 64, 91-104.
Davids, A. (1968). Psychological characteristics of high school male and female potential scientists in
comparison with academic underachievers. Psychology in the Schools, 3, 79-87.
Delias, M., & Gaier, E. L. (1970). Identification of creativity: The individual. Psychological Bulletin, 73,
55-73-

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
2Q2 G. J. F E I S T

Digman, J. M. (1990). Personality structure: Emergence of the five-factor model. Annual Review of Psy-
chology, 41, 417-440-
Dollinger, S. J., & Clancy, S. M. (1993). Identity, self, and personality. Part 2, Glimpses through the
autophotographic eye. Journal of Personality and Social Psychology, 64, 1064-1071.
Domino, G. (1974). Assessment of cinematographic creativity, journal of Personality and Social Psy-
chology, 30, 150-154.
Drevdahl, J. E., & Cattell, R. B. (1958). Personality and creativity in artists and writers. Journal of Clin-
ical Psychology, 14, 107-111.
Dudek, S. (1968). Regression and creativity. Journal of Nervous and Mental Disease, 147, 535-546.
Dudek, S. Z., Berneche, R., Be'rube', H., & Royer, S. (1991). Personality determinants of the commit-
ment to the profession of art. Creativity Research Journal, 4, 367-389.
Dudek, S. Z., & Hall, W. B. (1991). Personality consistency: Eminent architects 25 years later. Creativ-
ity Research Journal, 4, 1213-231.
Eiduson, B. T. (1958). Artist and non-artist: A comparative study. Journal of Personality, 26, 13-28.
Eiduson, B. T. (1962). Scientists: Their psychological world. New York: Basic.
Erickson, C. O., Gantz, B. S., & Stephenson, R. W. (1990). Logical and construct validation of a short-
form biographical inventory predictor of scientific creativity. Proceedings, 78th Annual Convention,
APA, 151-152.
Eysenck, H. J. (1990). Biological dimensions of personality. In L. A. Pervin (Ed.), Handbook of person-
ality theory and research (pp. 244-276). New York: Guilford.
Eysenck, H. J. (1993). Creativity and personality: Suggestions for a theory. Psychological Inquiry, 4,
147-178.
Eysenck, H. J. (1994). Creativity and personality: Word association, origence, and psychoticism. Cre-
ativity Research Journal, 7, 209-216.
Eysenck, H. J. (1995). Genius: The natural history of creativity. Cambridge University Press.
Farmer, H. S. (1988). Predicting who our future scientists and mathematicians will be. Behavioral and
Brain Sciences, 11, 190-191.
Feist, G. J. (1989). [Creativity in art and science students]. Unpublished raw data.
Feist, G. J. (1991). Synthetic and analytic thought: Similarities and differences among art and science
students. Creativity Research Journal, 4, 145-155.
Feist, G. J. (1993). A structural model of scientific eminence. Psychological Science, 4, 366-371.
Feist, G. J. (1994). Personality and working style predictors of integrative complexity: A study of scien-
tists' thinking about research and teaching. Journal of Personality and Social Psychology, 67,
474-484.
Feist, G. J. (1995, October). Do hostile and arrogant scientists become eminent or are eminent scientists
likely to become hostile and arrogant. Paper presented at the annual conference of the Society for
Social Studies of Science, Charlottesville, VA.
Feist, G. J. (in press). Affective states and traits in creativity: Evidence for non-linear relationships. In
M. A. Runco (Ed.), Creativity Research Handbook (Vol. 2.). Cresskill, NJ: Hampton.
Feist, G. J., & Barron, F. (1996). [Longitudinal study of 1950 graduate students]. Unpublished raw data.
Feist, G. J., & Gorman, M. E. (1998). The psychology of science: Review and integration of a nascent
discipline. Review of General Psychology, 2, 3-47.
Fodor, E. M. (1995). Subclinical manifestations of psychosis-proneness, ego-strength, and creativity.
Personality and Individual Differences, 18, 635-642.
Gantz, B. S, Erickson, C. O., & Stephenson, R. W. (1972). Some determinants of promotion in a
research and development population. Conference proceedings for the 72nd Annual Convention of
the American Psychological Association. Washington, DC: American Psychological Association.
Gardner, H. (1973). The arts and human development: A psychological study of the artistic process.
New York: Wiley.
Garwood, D. S. (1964). Personality factors related to creativity in young scientists. Journal of Abnormal
and Social Psychology, 68, 413-419.
Getzels, J. W. (1987). Creativity, intelligence, and problem finding: Retrospect and prospect. In S. G.
Isaksen (Ed.), Frontiers of creativity research (pp. 88-102). Buffalo, NY: Bearly.
Getzels, J. W, & Csikszentmihalyi, M. (1976). The creative vision. New York: Wiley.
Ghiselin, B. (Ed.). (1952). The creative process. New York: Mentor.
Gotz, K. O., & Gotz, K. (1979). Personality characteristics of professional artists. Perceptual and Motor
Skills, 49, 327-334.
Gough, H. G. (1961, February). A personality sketch of the creative research scientist. Paper presented
at the Fifth Annual Conference on Personnel and Industrial Relations Research, UCLA, Los Ange-
les, CA.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 293
Gough, H. G. (1976). What happens to creative medical students? Journal of Medical Education, 52,
461-467.
Gough, H. G. (1987). California Psychological Inventory: Administrators guide. Palo Alto, CA: Con-
sulting Psychologists Press.
Guastello, S., & Shissler, J. (1994). A two-factor taxonomy of creative behavior. Journal of Creative
Behavior, 28, 211-221.
Guilford, J. P. (1950). Creativity. American Psychologist, 5, 444-454.
Guilford, J. P. (1959). Traits of creativity. In H. H. Anderson (Ed.), Creativity and its cultivation (pp.
142-161). New York: Harper.
Guilford, J. P. (1987). A review of a quarter century of progress. In S. G. Isaksen (Ed.), Frontiers of cre-
ativity research (pp. 45-61). Buffalo, NY: Bearly.
Hall, W. B., & MacKinnon, D. W. (1969). Personality inventory correlates of creativity among architects.
Journal of Applied Psychology, 53, 322-326.
Ham, S., & Shaughnessy, M. F. (1992). Personality and scientific promise. Psychological Reports, 70,
971-975-
Hammer, E. F. (1966). Personality patterns in young creative artists. Adolescence, 1, 327-350.
Hammond, J., & Edelmann, R. J. (1991). The act of being: Personality characteristics of professional
actors, amateur actors and non-actors. In G. Wilson (Ed.), Psychology and performing arts (pp.
123-131). Amsterdam: Swets & Zeitlinger.
Helmreich, R. L., Spence, J. T, Beane, W. E., Lucker, G. W, & Matthews, K. A. (1980). Making it in
academic psychology: Demographic and personality correlates of attainment. Journal of Personality
and Social Psychology, 39, 896-908.
Helmreich, R. L., Spence, J. T, & Pred, R. S. (1988). Making it without losing it: Type A, achievement
motivation and scientific attainment revisited. Personality and Social Psychology Bulletin, 14,
495-5O4-
Helson, R. (1971). Women mathematicians and the creative personality. Journal of Consulting and Clin-
ical Psychology, 36, 210-220.
Helson, R. (1977). The creative spectrum of authors of fantasy. Journal of Personality, 4$, 310-326.
Helson, R. (1987). Which of those young women with creative potential became productive? Part 2,
From college to midlife. In R. Hogan, & W. H. Jones (Eds.), Perspectives in personality (Vol. 2, pp.
51-92). Greenwich, CN: JAI.
Helson, R., & Crutchfield, R. S. (1970). Mathematicians: The creative researcher and the average Ph.D.
Journal of Consulting and Clinical Psychology, 34, 250-257.
Helson, R., Roberts, B., & Agronick, G. (1995). Enduringness and change in creative personality and
prediction of occupational creativity. Journal of Personality and Social Psychology, 69, 1173-1183.
Holahan, C. K., & Sears, R. R. (1995). The gifted group in later maturity. Stanford, CA: Stanford Uni-
versity Press.
Holland, J. (i960). The prediction of college grades from personality and aptitude variables. Journal of
Educational Psychology, $1, 245-254.
Holland, J. L., & Baird, L. L. (1968). The preconscious activity scale: The development and validation
of an originality measure. Journal of Creative Behavior, 2, 217-225.
Hudson, L. (1958). Undergraduate academic record of Fellows of the Royal Society. Nature, 182, 1326.
Ikpaahindi, L. (1987). The relationship between the needs for achievement, affiliation, power, and sci-
entific productivity among Nigerian veterinary surgeons. Journal of Social Psychology, 12J, 535-537.
Isen, A., Daubman, K. A., & Nowicki, G. P. (1987). Positive affect facilitates creative problem solving.
Journal of Personality and Social Psychology, 52, 1122-1131.
Jamison, K. R. (1993). Touched with fire: Manic-depressive illness and the artistic temperament. New
York: Free Press.
John, O. P. (1990). The "bigfive"factor taxonomy: Dimensions of personality in the natural language
and in questionnaires. In L. A. Pervin (Ed.), Handbook of personality research and theory (pp.
66-100). New York: Guilford.
Kemp, A. (1981). The personality structure of the musician. Part I, Identifying a profile of traits for the
performer. Psychology of Music, 9, 3-14.
Kline, P., & Lapham, S. L. (1992). Personality and faculty in British universities. Personality and Indi-
vidual Differences, 13, 855-857.
Kuhn, T. S. (1970). The structure of scientific revolutions (2nd ed.). Chicago: University of Chicago
Press.
Lacey, L. A., & Erickson, C. E. (1974). Psychology of the scientist. Part 31, Discriminability of a cre-
ativity scale for the Adjective Check List among scientists and engineers. Psychological Reports, 34,
8

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
G
294 « J- F E I S T
Lindsay, B. (1978). Leadership giftedness: Developing a profile. Journalfor the Education of the Gifted,
1, 63-69-
Lubinski, D., & Benbow, C. P. (1994). The study of mathematically precocious youth: The first three
decades of a planned 50-year study of intellectual talent. In R. F. Subotnik & K. D. Arnold (Eds.),
Beyond Terrnan: Contemporary longitudinal studies of giftedness and talent (pp. 255-281). Nor-
wood, NJ: Ablex.
Ludwig, A. M. (1995). The price of greatness. New York: Guilford.
MacKinnon, D. W. (i960). The highly effective individual. Teachers College Record, 61, 367-378.
MacKinnon, D. W. (1962). The nature and nurture of creative talent. American Psychologist, 17,
484-495.
MacKinnon, D. W. (1970). Creativity: A multi-faceted phenomenon. In J. Roslanksy (Ed.), Creativity
(pp. 19-32). Amsterdam: North-Holland.
MacKinnon, D. W. (1978). In search of human effectiveness. Buffalo, NY: Bearly.
Magnusson, D., & Backteman, G. (1978). Longitudinal stability of person characteristics: Intelligence
and creativity. Applied Psychological Measurement, 2, 481-490.
Mansfield, R. S., & Busse, T. V. (1981). The psychology of creativity and discovery: Scientists and their
work. Chicago: Nelson-Hall.
Marchant-Haycox, S. E., & Wilson, G. D. (1992). Personality and stress in performing artists. Person-
ality and Individual Differences, 13, 1061-1068.
Marland, S. P., Jr. (1972). Education of the gifted and talented. Washington, DC: U.S. Government
Printing Office.
Martindale, C. (1975). Romantic progression: The psychology of literary history. Washington, DC:
Hemisphere.
Martindale, C. (1981). Cognition and consciousness. Homewood, IL: Dorsey.
Martindale, C , & Armstrong, J. (1974). The relationship of creativity to cortical activation and its oper-
ant control. Journal of Genetic Psychology, 124, 311-320.
Martindale, C , & Greenough, J. (1973). The differential effect of increased arousal on creative and
intellectual performance. Journal of Genetic Psychology, 123, 329-335.
Martindale, C , & Hasenfus, N. (1978). EEG differences as a function of creativity, stage of the creative
process and effort to be original. Biological Psychology, 6, 157-167.
Martindale, C , Hines, D., Mitchell, L., & Covello, E. (1984). EEG alpha asymmetry and creativity. Per-
sonality and Individual Differences, 5, 77-86.
McCrae, R. R. (1987). Creativity, divergent thinking, and openness to experience. Journal of Personal-
ity and Social Psychology, 52, 1258-1265.
McCrae, R. R., & John, O. P. (1992). An introduction to the five-factor model and its applications. Jour-
nal of Personality, 60, 175-215.
McDermid, C. D. (1965). Some correlates of creativity in engineering personnel. Journal of Applied
Psychology, 49, 14-19.
Mendelsohn. G. A. (1976). Associative and attentional processes in creative performance. Journal of
Personality, 44, 341-369.
Milgram, R. M., & Hong, E. (1994). Creative thinking and creative performance in adolescents as pre-
dictors of creative attainments in adults: A follow-up study after 18 years. In R. F. Subotnik & K. D.
Arnold (Eds.), Beyond Terrnan: Contemporary longitudinal studies of giftedness and talent (pp.
212-228). Norwood, NJ: Ablex.
Mohan, J., & Tiwana, M. (1987). Personality and alienation of creative writers: A brief report. Person-
ality and Individual Differences, 8, 449.
Mumford, M. D., Costanza, D. P., Threlfall, K. V, Baughman, W. A., & Reiter-Palmon, R. (1993). Per-
sonality variables and problem-construction activities: An exploratory investigation. Creativity
Research Journal, 6, 365-389.
Mumford, M. D., & Gustafson, S. B. (1988). Creativity syndrome: Integration, application, and innova-
tion. Psychological Bulletin, 103, 27-43.
Ochse, R. (1990). Before the gates of excellence: The determinants of creative genius. Cambridge Uni-
versity Press.
Parloff, M. B., & Datta, L. (1965). Personality characteristics of the potentially creative scientist. Sci-
ence and Psychoanalysis, 8, 91-105.
Parloff, M. B., Datta, L., Kleman, M., & Handlon, J. H. (1968). Personality characteristics which dif-
ferentiate creative male adolescents and adults. Journal of Personality, 36, 528-552.
Pufal-Struzik, I. (1992). Differences in personality and self-knowledge of creative persons at different
ages: A comparative analysis. Special Issue: Geragogics: European research in gerontological educa-
tion. Gerontology ir Geriatrics Education, 13, 71-90.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
Influence of Personality on Artistic and Scientific Creativity 295
Pyryt, M. C. (1992). The fulfillment of promise revisited: A discriminant analysis of factors predicting
success in the Terman study. Roeper Review, 15, 178-179.
Richards, R. L. (1994). Creativity and bipolar mood swings: Why the association? In M. P. Shaw &
M. A. Runco (Eds.), Creativity and affect (pp. 44-72). Norwood, NJ: Ablex.
Richards, R. L., & Kinney, D. K. (1990). Mood swings and creativity. Creativity Research Journal, 3,
202-217.
Richards, R. L., Kinney, D. K., Lunde, I., Benet, M., & Merzel, A. (1988). Creativity in manic-
depressives, cyclothymes, their normal relatives, and control subjects. Journal of Abnormal Psychol-
ogy, Q7, 281-289.
Roco, M. (1993). Creative personalities about creative personality in science. Revue Roumaine de Psy-
chologie, 37, 27-36.
Roe, A. (1952). The making of a scientist. New York: Dodd, Mead.
Roe, A. (1953). A psychological study of eminent psychologists and anthropologists, and a comparison
with biological and physical scientists. Psychological Monographs: General and Applied, 67, 1-55.
Rosenthal, R., & Rosnow, R. L. (1991). Essentiah of behavioral research: Methods and data analysis
(2nded.). New York: McGraw-Hill.
Rossman, B. B., & Horn, J. L. (1972). Cognitive, motivational and temperamental indicants of creativ-
ity and intelligence. Journal of Educational Measurement, 9, 265-286.
Rothenberg, A. (1990). Creativity and madness: New findings and old stereotypes. Baltimore: Johns
Hopkins University Press.
Rothenberg, A., & Hausman, C. R. (Eds.). (1976). The creativity question. Durham, NC: Duke Uni-
versity Press.
Runco, M. A., & Bahleda, M. D. (1986). Implicit theories of artistic, scientific, and everyday creativity.
Journal of Creative Behavior, 20, 93-98.
Rushton, J. P. (1990). Creativity, intelligence, and psychoticism. Personality and Individual Differences,
12, 1291-1298.
Rushton, J. P., Murray, H. G., & Paunonen, S. V. (1983). Personality, research creativity, and teaching
effectiveness in university professors. Scientometrics, 5, 93-116.
Rushton, J. P., Murray, H. G., & Paunonen, S. V. (1987). Personality characteristics associated with high
research productivity. In D. Jackson & J. P. Rushton (Eds.), Scientific excellence (pp. 129-148). Bev-
erly Hills, CA: Sage.
Russ, S. (1993). Affect and creativity: The role of affect and play in the creative process. Hillsdale, NJ:
Erlbaum.
Schaefer, C. E. (1969). The self-concept of creative adolescents. Journal of Psychology, 72, 233-242.
Schaefer, C. E. (1973). A five-year follow-up study of the self-concept of creative adolescents. Journal
of Genetic Psychology, 123, 163-170.
Shapiro, R. J. (1968). Creative research scientists. Psychologia Africana Monograph Supplement,
4(180).
Sheldon, K. M. (1995). Creativity and self-determination in personality. Creativity Research Journal, 8,
23-36-
Shelton, J., & Harris, T. L. (1979). Personality characteristics of art students. Psychological Reports, 44,
949-950-
Simon, H. (1974). The work habits of eminent scientists. Sociology of Work and Occupations, 1,
327-335.
Simonton, D. K. (1988). Scientific genius: A psychology of science. Cambridge University Press.
Simonton, D. K. (1989). Chance-configuration theory of scientific creativity. In B. Gholson, W. R.
Shadish, R. A. Neimeyer, & A. C. Houts (Eds.), Psychology of science, (pp. 170-213). Cambridge
University Press.
Smithers, A. G., & Batcock, A. (1970). Success and failure among social scientists and health scientists
at a technological university. British Journal of Educational Psychology, 40, 144-153.
Stein, M. (1968). Creativity. In E. F. Borgatta & W. W. Lambert (Eds.), Handbook of personality theory
and research (pp. 900-942). Chicago: Rand McNally.
Sternberg, R. J. (1986). The triarchic mind: A new theory of human intelligence. New York: Viking.
Sternberg, R. J. (1988). A three-facet model of creativity. In R. J. Sternberg (Ed.), The nature of cre-
ativity (pp. 125-147). Cambridge University Press.
Sternberg, R. J., & Lubart, T. (1995). Defying the crowd. New York: Free Press.
Sternberg, R. J., & Lubart, T. (1996). Investing in creativity. American Psychologist, 51, 677-688.
Stohs, J. M. (1990). Young adult predictors and midlife outcomes of male fine art careers. Career Devel-
opment Quarterly, 38, 213-229.
Storr, A. (1988). Solitude: A return to the self New York: Free Press.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016
296 G. J. FEIST
Subotnik, R. R, & Arnold, K. D. (Eds.). (1994). Beyond Terman: Contemporary longitudinal studies of
giftedness and talent. Norwood, NJ: Ablex.
Subotnik, R. R, Duschl, R. A., & Selmon, E. H. (1993). Retention and attrition of science talent: A lon-
gitudinal study of Westinghouse science talent search winners. International Journal of Science Edu-
cation, 15, 61-72.
Subotnik, R. R, & Steiner, C. L. (1992). Adult manifestations of adolescent talent in science. Roeper
Review, 15, 164-169.
Terman, L. M. (1954). Scientists and nonscientists in a group of 800 men. Psychological Monographs,
68, Whole No. 378.
Terman, L. M. (1955). Are scientists different? Scientific American, 192, 25-29.
Tomlinson-Keasey, C , & Keasey, C. B. (1993). Graduating from college in the 1930s: Terman genetic
studies of genius. In K. D. Hulbert & S. D. Schuster (Eds.), Women's lives through time: Educated
women of the twentieth century (pp. 63-92). San Francisco: Jossey-Bass.
Tomlinson-Keasey, C., & Little, T. D. (1990). Predicting educational attainment, occupational achieve-
ment, intellectual skill, and personal adjustment among gifted men and women. Journal of Educa-
tional Psychology, 82, 442-455.
Trost, G. (1993). Prediction of excellence in school, university and work. In K. A. Heller, P. J. Monks, &
A. H. Passow (Eds.), International handbook of research and development of giftedness and talent
(pp. 325-336). Oxford: Pergamon.
Van Zelst, R. H., & Kerr, W. A. (1954). Personality self-assessment of scientific and technical personnel.
Journal of Applied Psychology, 38, 145-147.
Vernon, P. E. (Ed.). (1970). Creativity. Harmondsworth: Penguin.
Walker, A. M., Koestner, R., & Hum, A. (1995). Personality correlates of depressive style in autobi-
ographies of creative achievers. Journal of Creative Behavior, 2Q, 75-94.
Wallach, M. A. (1970). Creativity. In P. H. Mussen (Ed.), Manual of child psychology (pp. 1211-1272).
New York: Wiley.
Wills, G. I. (1983). A personality study of musicians working in the popular field. Personality and Indi-
vidual Differences, 5, 359-360.
Wilson, G. D. (1984). The personality of opera singers. Personality and Individual Differences, 5,
195-201.
Wilson, G. D., & Jackson, C. (1994). The personality of physicists. Personality and Individual Differ-
ences, 16, 187-189.
Winner, E., & Martino, G. (1993). Giftedness in the visual arts and music. In K. A. Heller, R J. Monks,
& A. H. Passow (Eds.), International handbook of research and development of giftedness and talent
(pp. 253-281). Oxford: Pergamon.
Wise, L. L., Steel, L., & McDonald, C. (1979). Origins and career consequences of sex differences in
high school mathematics achievement. Washington D.C: American Institute for Research.
Wispe, L. G. (1963). Traits of eminent American psychologists. Science, 141, 1256-1261.
Woody, E., & Claridge, G. (1977). Psychoticism and thinking. British Journal of Social and Clinical Psy-
chology, 16, 241-248.
Zeldow, P. B. (1973). Replication and extension of the personality profile of "artists in the making." Psy-
chological Reports, 33, 541-542.

Downloaded from https:/www.cambridge.org/core. University of Sussex Library, on 12 Mar 2017 at 13:49:56, subject to the Cambridge Core
terms of use, available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9780511807916.016

You might also like