首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
The dominant view among philosophers of perception is that color experiences, like color judgments, are essentially representational: as part of their very nature color experiences possess representational contents which are either accurate or inaccurate. My starting point in assessing this view is Sydney Shoemaker’s familiar account of color perception. After providing a sympathetic reconstruction of his account, I show how plausible assumptions at the heart of Shoemaker’s theory make trouble for his claim that color experiences represent the colors of things. I consider various ways of trying to avoid the objection, and find all of the responses wanting. My conclusion is that we have reason to be skeptical of the orthodox view that color experiences are constitutively representational.  相似文献   

3.
What is color vision?   总被引:3,自引:0,他引:3  
  相似文献   

4.
A possible source for the advantage of location cueing over non-spatial cueing is that orienting attention by a location cue is feasible prior to stimulus onset, whereas that is normally not the case with orienting by a non-spatial cue. To examine how critical that source is for observing an advantage, we eliminated it: In a color-preview condition, subjects were to detect a target presented on the background of one of two differently colored circles (where color-location assignment was random). In a no-preview condition, the circles were both gray, but the target was either red or green (where color assignment was random). Cue type (location vs color) was also manipulated. The color preview in Experiment 1 (in which color onset preceded cue onset) was found helpful: Whereas a substantial disparity in validity effects of the two cue types was obtained with no preview, no significant difference was found when a color preview was introduced. The validity effects of both cue types were found to be about the same also in Experiment 2, in which color onset was exactly synchronized with cue onset, and SOA was manipulated. Furthermore, the absence of an SOA x cue type interaction indicated that the time course of the color cue validity did not lag after the time course of the location cue validity, which seems incompatible with the hypothesis that a color cue cannot affect orienting without first computing a location from it prior to cue onset. Overall, the results suggest that the time course of color cueing is not inherently different from that of location cueing once its main disadvantages are removed.  相似文献   

5.
Four experiments are reported in which the effects of peripheral cues on visual orienting were investigated. In the luminance condition, the cues consisted of a peripheral change in stimulus luminance. In the isoluminance condition, the cues consisted of an isoluminant color change, using the transient tritanopic technique. In Experiments 1 and 2, it was found that peripheral luminance cues captured attention, whereas peripheral isoluminance cues did not. In Experiments 3 and 4, the participants detected a peripheral target that was also isoluminant with the background. Under these conditions, it was found that both luminance and isoluminance cues captured attention. The results are discussed in terms of the roles of the dorsal and ventral streams in visual orienting, and it is concluded that our findings provide partial support for the contingent involuntary orienting hypothesis of C. Folk and colleagues.  相似文献   

6.
Typically, the search for order in grapheme–color synesthesia has been conducted by looking at the frequency of certain letter–color associations. Here, we report stronger associations when second-order similarity mappings are examined—specifically, mappings between the synesthetic colors of letters and letter shape, frequency, and position in the alphabet. The analyses demonstrate that these relations are independent of one other. More strikingly, our analyses show that each of the letter–color mappings is restricted to one dimension of color, with letter shape and ordinality linked to hue, and letter frequency linked to luminance. These results imply that synesthetic associations are acquired as the alphabet is learned, with associations involving letter shape, ordinality, and frequency being made independently and idiosyncratically. Because these mappings of similarity structure between domains (letters and colors) are similar to those found in numerous other cognitive and perceptual domains, they imply that synesthetic associations operate on principles common to many aspects of human cognition.  相似文献   

7.
Roberson and Davidoff (2000) found that color categorical perception (CP; better cross-category than within-category discrimination) was eliminated by verbal, but not by visual, interference presented during the interstimulus interval (ISI) of a discrimination task. On the basis of this finding, Roberson and Davidoff concluded that CP was mediated by verbal labels, and not by perceptual mechanisms, as is generally assumed. Experiment 1 replicated their results. However, it was found that if the interference type was uncertain on each trial (Experiment 2), CP then survived verbal interference. Moreover, it was found that the target color name could be retained across the ISI even with verbal interference (Experiment 3). We therefore conclude that color CP may indeed involve verbal labeling but that verbal interference does not necessarily prevent it.  相似文献   

8.
In the non-color-word Stroop task, university students' response latencies were longer for low-frequency than for higher frequency target words. Visual identity primes facilitated color naming in groups reading the prime silently or processing it semantically (Experiment 1) but did not when participants generated a rhyme of the prime (Experiment 3). With auditory identity primes, generating an associate or a rhyme of the prime produced interference (Experiments 2 and 3). Color-naming latencies were longer for nonwords than for words (Experiment 4). There was a small long-term repetition benefit in color naming for low-frequency words that had been presented in the lexical decision task (Experiment 5). Facilitation of word recognition speeds color naming except when phonological activation of the base word increases response competition.  相似文献   

9.
In an array of elements whose colors vary can we selectively choose to process all the items of a particular color preferentially in relation to those of another color? We addressed this question by presenting subjects with arrays containing many elements, and recording reaction times to a luminance change of one of the elements. Half the elements had one color and the other half another color--the spatial distribution being random. In two tasks--a simple detection of this change or a choice reaction time to the polarity of the change--we found that reaction times were independent of the number of items in the array. Cuing the subjects as to the color of the target item had no significant influence on the detection task, but subjects were faster if cued for the discrimination task. A further experiment replicated these findings and examined possible costs and benefits. Our final experiment separated the roles of attentional guidance and postattentional processes by having subjects judge the orientation of the target element and varying the magnitude of the target flash that defined which element was the target. We found that this judgment was also affected by color cuing, and that the size of the effect interacted with the flash strength, suggesting that color cuing has its influence at the stage of attentional guidance. We conclude that subjects can selectively attend to items on the basis of color given the appropriate task and stimulus dynamics.  相似文献   

10.
We investigated the physiological mechanism of grapheme–color synesthesia using metacontrast masking. A metacontrast target is rendered invisible by a mask that is delayed by about 60 ms; the target and mask do not overlap in space or time. Little masking occurs, however, if the target and mask are simultaneous. This effect must be cortical, because it can be obtained dichoptically. To compare the data for synesthetes and controls, we developed a metacontrast design in which nonsynesthete controls showed weaker dichromatic masking (i.e., the target and mask were in different colors) than monochromatic masking. We accomplished this with an equiluminant target, mask, and background for each observer. If synesthetic color affected metacontrast, synesthetes should show monochromatic masking more similar to the weak dichromatic masking among controls, because synesthetes could add their synesthetic color to the monochromatic condition. The target–mask pairs used for each synesthete were graphemes that elicited strong synesthetic colors. We found stronger monochromatic than dichromatic U-shaped metacontrast for both synesthetes and controls, with optimal masking at an asynchrony of 66 ms. The difference in performance between the monochromatic and dichromatic conditions in the synesthetes indicates that synesthesia occurs at a later processing stage than does metacontrast masking.  相似文献   

11.
Mandik understands color-consciousness conceptualism to be the view that one deploys in a conscious qualitative state concepts for every color consciously discriminated by that state. Some argue that the experimental evidence that we can consciously discriminate barely distinct hues that are presented together but cannot do so when those hues are presented in short succession suggests that we can consciously discriminate colors that we do not conceptualize. Mandik maintains, however, that this evidence is consistent with our deploying a variety of nondemonstrative concepts for those colors and so does not pose a threat to conceptualism. But even if Mandik has shown that we deploy such concepts in these experimental conditions, there are cases of conscious states that discriminate colors but do not involve concepts of those colors. Mandik's arguments sustain only a theory in the vicinity of conceptualism: The view that we possess concepts for every color we can discriminate consciously, but need not deploy those concepts in every conscious act of color discrimination.  相似文献   

12.
Naor-Raz G  Tarr MJ  Kersten D 《Perception》2003,32(6):667-680
The role of color in object representation was examined by using a variation of the Stroop paradigm in which observers named the displayed colors of objects or words. In experiment 1, colors of color-diagnostic objects were manipulated to be either typical or atypical of the object (eg a yellow banana versus a purple banana). A Stroop-like effect was obtained, with faster color-naming times for the typical as compared to the atypical condition. In experiment 2, naming colors on words specifying these same color-diagnostic objects reversed this pattern, with the typical condition producing longer response times than the atypical condition. In experiment 3, a blocked condition design that used the same words and colors as experiment 2 produced the standard Stroop-like facilitation for the typical condition. These results indicate that color is an intrinsic property of an object's representation at multiple levels. In experiment 4, we examined the specific level(s) at which color-shape associations arise by following the tasks used in experiments 1 and 2 with a lexical-decision task in which some items were conceptually related to items shown during color naming (eg banana/monkey). Priming for these associates was observed following color naming of words, but not pictures, providing further evidence that the color-shape associations responsible for the differing effects obtained in experiments 1 and 2 are due to the automatic activation of color-shape associations at different levels of representation.  相似文献   

13.
In the 1900’s, L. A. Orbeli, on I. P. Pavlov’s request, attempted to establish color discrimination in dogs. Previous experiments used the motor discrimination method and gave inconsistent results. In contrast, Orbeli used the salivary conditional reflex method, which he considered to be more precise than the method that relied on erratic movements of a dog. After experimentation that lasted about one and a half years, Orbeli failed to establish color discrimination. When subsequent experiments by Russians and Germans yielded positive results with the motor discrimination method, Orbeli switched to this method and also obtained positive results. These findings were confirmed by most subsequent experiments on color discrimination in dogs. The utility of Pavlovian conditioning in sensory experimentation was not universally justified and its importance was greater in the study of learning.  相似文献   

14.
Much work has been done on visual discrimination in primates over the past decade. In contrast, very little is known about the relevance of non-visual information in discrimination learning. We investigated weight and achromatic color (color, henceforth) discrimination in bonobos, gorillas and orangutans, using the exchange paradigm in which subjects have to give objects to the experimenter in order to receive a reward. Unlike previous studies, subjects were not trained to lift objects because lifting the objects was an integral part of the exchange procedure. This methodology also allowed us a direct comparison between visual and weight discrimination. We presented 12 subjects (5 bonobos, 2 gorillas and 5 orangutans) with two sets of objects corresponding to two conditions. The objects in the color condition (white/black) differed only in color and those in the weight condition (light/heavy) differed only in weight. Five apes learned to discriminate weight and six to discriminate color. Subjects learned color discrimination faster than weight discrimination. Our results suggest that bonobos and orangutans are sensitive to differences in weight and able to learn discriminating objects that differ in this property.  相似文献   

15.
In three experiments, we tested a relative-speed-of-processing account of color–word contingency learning, a phenomenon in which color identification responses to high-contingency stimuli (words that appear most often in particular colors) are faster than those to low-contingency stimuli. Experiment 1 showed equally large contingency-learning effects whether responding was to the colors or to the words, likely due to slow responding to both dimensions because of the unfamiliar mapping required by the key press responses. For Experiment 2, participants switched to vocal responding, in which reading words is considerably faster than naming colors, and we obtained a contingency-learning effect only for color naming, the slower dimension. In Experiment 3, previewing the color information resulted in a reduced contingency-learning effect for color naming, but it enhanced the contingency-learning effect for word reading. These results are all consistent with contingency learning influencing performance only when the nominally irrelevant feature is faster to process than the relevant feature, and therefore are entirely in accord with a relative-speed-of-processing explanation.  相似文献   

16.
Research suggests that language comprehenders simulate visual features such as shape during language comprehension. In sentence-picture verification tasks, whenever pictures match the shape or orientation implied by the previous sentence, responses are faster than when the pictures mismatch implied visual aspects. However, mixed results have been demonstrated when the sentence-picture paradigm was applied to color (Connell, Cognition, 102(3), 476–485, 2007; Zwaan & Pecher, PLOS ONE, 7(12), e51382, 2012). One of the aims of the current investigation was to resolve this issue. This was accomplished by conceptually replicating the original study on color, using the same paradigm but a different stimulus set. The second goal of this study was to assess how much perceptual information is included in a mental simulation. We examined this by reducing color saturation, a manipulation that does not sacrifice object identifiability. If reduction of one aspect of color does not alter the match effect, it would suggest that not all perceptual information is relevant for a mental simulation. Our results did not support this: We found a match advantage when objects were shown at normal levels of saturation, but this match advantage disappeared when saturation was reduced, yet still aided in object recognition compared to when color was entirely removed. Taken together, these results clearly show a strong match effect for color, and the perceptual richness of mental simulations during language comprehension.  相似文献   

17.
Processing colored pictures of objects results in a preference to choose the former color for a specific object in a subsequent color choice test (Wippich & Mecklenbr?uker, 1998). We tested whether this implicit memory effect is independent of performances in episodic color recollection (recognition). In the study phase of Experiment 1, the color of line drawings was either named or its appropriateness was judged. We found only weak implicit memory effects for categorical color information. In Experiment 2, silhouettes were colored by subjects during the study phase. Performances in both the implicit and the explicit test were good. Selections of "old" colors in the implicit test, though, were almost completely confined to items for which the color was also remembered explicitly. In Experiment 3, we applied the opposition technique in order to check whether we could find any implicit effects regarding items for which no explicit color recollection was possible. This was not the case. We therefore draw the conclusion that implicit color preference effects are not independent of explicit recollection, and that they are probably based on the same episodic memory traces that are used in explicit tests.  相似文献   

18.
19.
Why just turquoise? Remarks on the evolution of color terms   总被引:1,自引:0,他引:1  
Summary The location of the foci of green and blue in the perceptual color solid indicates that there is space for a derived color term between these two hues. Diachronic and synchronic linguistic studies on color term lexica explain that a term for turquoise is likely to develop into a derived basic color term (in the Kay and McDaniel definition), at present in languages of industrialized countries. In addition to the hypothesis of Zimmer (1982) that türkis (in German) is the result of universal production system for color terms, cultural, social, and psychological factors influence the evolution of new basic color terms.  相似文献   

20.
Young children experience difficulties establishing conceptual representations of color compared with everyday objects. We argue that comparing the development of color cognition to that of familiar objects is inappropriate since color is a perceptual attribute that can be abstracted from an object and by itself lacks functional significance. Instead, we compared the recognition, perceptual saliency, and naming of color to that of three other perceptual object attributes (motion, form, and size) in 47 children aged 2 to 5 years as a function of language age. Results revealed that, although color was perceptually salient relative to the other visual attributes, no selective impairment to color cognition (recognition and naming) was found relative to the three other visual attributes tested. Thus, when the appropriate comparisons are made, we find no special delay in the development of color conceptualization. Furthermore, the striking disparity between perceptual saliency and cognition of color in our youngest age groups suggests that perceptual saliency has little influence on the conceptual development of color.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号