nips nips2011 nips2011-232 knowledge-graph by maker-knowledge-mining
Source: pdf
Author: Vikas C. Raykar, Shipeng Yu
Abstract: With the advent of crowdsourcing services it has become quite cheap and reasonably effective to get a dataset labeled by multiple annotators in a short amount of time. Various methods have been proposed to estimate the consensus labels by correcting for the bias of annotators with different kinds of expertise. Often we have low quality annotators or spammers–annotators who assign labels randomly (e.g., without actually looking at the instance). Spammers can make the cost of acquiring labels very expensive and can potentially degrade the quality of the consensus labels. In this paper we formalize the notion of a spammer and define a score which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a high score close to one. 1 Spammers in crowdsourced labeling tasks Annotating an unlabeled dataset is one of the bottlenecks in using supervised learning to build good predictive models. Getting a dataset labeled by experts can be expensive and time consuming. With the advent of crowdsourcing services (Amazon’s Mechanical Turk being a prime example) it has become quite easy and inexpensive to acquire labels from a large number of annotators in a short amount of time (see [8], [10], and [11] for some computer vision and natural language processing case studies). One drawback of most crowdsourcing services is that we do not have tight control over the quality of the annotators. The annotators can come from a diverse pool including genuine experts, novices, biased annotators, malicious annotators, and spammers. Hence in order to get good quality labels requestors typically get each instance labeled by multiple annotators and these multiple annotations are then consolidated either using a simple majority voting or more sophisticated methods that model and correct for the annotator biases [3, 9, 6, 7, 14] and/or task complexity [2, 13, 12]. In this paper we are interested in ranking annotators based on how spammer like each annotator is. In our context a spammer is a low quality annotator who assigns random labels (maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator). Spammers can significantly increase the cost of acquiring annotations (since they need to be paid) and at the same time decrease the accuracy of the final consensus labels. A mechanism to detect and eliminate spammers is a desirable feature for any crowdsourcing market place. For example one can give monetary bonuses to good annotators and deny payments to spammers. The main contribution of this paper is to formalize the notion of a spammer for binary, categorical, and ordinal labeling tasks. More specifically we define a scalar metric which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a score close to one (see Figure 4). We summarize the multiple parameters corresponding to each annotator into a single score indicative of how spammer like the annotator is. While this spammer score was implicit for binary labels in earlier works [3, 9, 2, 6] the extension to categorical and ordinal labels is novel and is quite different from the accuracy computed from the confusion rate matrix. An attempt to quantify the quality of the workers based on the confusion matrix was recently made by [4] where they transformed the observed labels into posterior soft labels based on the estimated confusion 1 matrix. While we obtain somewhat similar annotator rankings, we differ from this work in that our score is directly defined in terms of the annotator parameters (see § 5 for more details). The rest of the paper is organized as follows. For ease of exposition we start with binary labels (§ 2) and later extend it to categorical (§ 3) and ordinal labels (§ 4). We first specify the annotator model used, formalize the notion of a spammer, and propose an appropriate score in terms of the annotator model parameters. We do not dwell too much on the estimation of the annotator model parameters. These parameters can either be estimated directly using known gold standard 1 or the iterative algorithms that estimate the annotator model parameters without actually knowing the gold standard [3, 9, 2, 6, 7]. In the experimental section (§ 6) we obtain rankings for the annotators using the proposed spammer scores on some publicly available data from different domains. 2 Spammer score for crowdsourced binary labels j Annotator model Let yi ∈ {0, 1} be the label assigned to the ith instance by the j th annotator, and let yi ∈ {0, 1} be the actual (unobserved) binary label. We model the accuracy of the annotator separately on the positive and the negative examples. If the true label is one, the sensitivity (true positive rate) αj for the j th annotator is defined as the probability that the annotator labels it as one. j αj := Pr[yi = 1|yi = 1]. On the other hand, if the true label is zero, the specificity (1−false positive rate) β j is defined as the probability that annotator labels it as zero. j β j := Pr[yi = 0|yi = 0]. Extensions of this basic model have been proposed to include item level difficulty [2, 13] and also to model the annotator performance based on the feature vector [14]. For simplicity we use the basic model proposed in [7] in our formulation. Based on many instances labeled by multiple annotators the maximum likelihood estimator for the annotator parameters (αj , β j ) and also the consensus ground truth (yi ) can be estimated iteratively [3, 7] via the Expectation Maximization (EM) algorithm. The EM algorithm iteratively establishes a particular gold standard (initialized via majority voting), measures the performance of the annotators given that gold standard (M-step), and refines the gold standard based on the performance measures (E-step). Who is a spammer? Intuitively, a spammer assigns labels randomly—maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator. More precisely an annotator is a spammer if the probability j of observed label yi being one given the true label yi is independent of the true label, i.e., j j Pr[yi = 1|yi ] = Pr[yi = 1]. This means that the annotator is assigning labels randomly by flipping a coin with bias without actually looking at the data. Equivalently (1) can be written as j j Pr[yi = 1|yi = 1] = Pr[yi = 1|yi = 0] which implies αj = 1 − β j . (1) j Pr[yi = 1] (2) Hence in the context of the annotator model defined earlier a perfect spammer is an annotator for whom αj + β j − 1 = 0. This corresponds to the diagonal line on the Receiver Operating Characteristic (ROC) plot (see Figure 1(a)) 2 . If αj + β j − 1 < 0 then the annotators lies below the diagonal line and is a malicious annotator who flips the labels. Note that a malicious annotator has discriminatory power if we can detect them and flip their labels. In fact the methods proposed in [3, 7] can automatically flip the labels for the malicious annotators. Hence we define the spammer score for an annotator as S j = (αj + β j − 1)2 (3) An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 1 One of the commonly used strategy to filter out spammers is to inject some items into the annotations with known labels. This is the strategy used by CrowdFlower (http://crowdflower.com/docs/gold). 2 Also note that (αj + β j )/2 is equal to the area shown in the plot and can be considered as a non-parametric approximation to the area under the ROC curve (AUC) based on one observed point. It is also equal to the Balanced Classification Rate (BCR). So a spammer can also be defined as having BCR or AUC equal to 0.5. 2 Equal accuracy contours (prevalence=0.5) 0.9 1 Good Annotators Biased Annotators 0.8 0.9 Sensitivity Sensitivity ( αj ) Spammers 0.5 0.4 j 0.3 j 0.6 0.5 0.4 4 0. 5 0. 3 0. 7 0. 6 0. 4 0. 5 0. 2 0. 3 0. Malicious Annotators 0.2 0.4 0.6 1−Specificity ( βj ) 0.8 0.1 1 0. 0. 2 0. 3 0. 1 0. 0.6 0.5 1 0. 2 0. 2 0. 1 0. 0.4 3 1 0. 0. 2 0. 4 0. 0.2 4 0. 5 0. 0 0 1 3 0. 0.3 0.2 Biased Annotators 4 0.7 6 0. 4 0. 0.8 7 0.3 Area = (α +β )/2 0.2 0 0 0.9 0. 8 0. 8 0. 0.7 6 0. .5 0 5 0. 0.7 [ 1−βj, αj ] 0.6 0.1 6 0. 0.8 0.7 Equal spammer score contours 1 7 0. 8 0. 9 0. Sensitivity 1 (a) Binary annotator model 0.1 1 0. 2 0. 3 0. 0.2 0.4 0.6 1−Specificity 0.8 1 1 0. 0 0 (b) Accuracy 0.2 3 0. 4 0. 0.4 0.6 1−Specificity 5 0. .6 7 0 0. 8 0. 0.8 1 (c) Spammer score Figure 1: (a) For binary labels an annotator is modeled by his/her sensitivity and specificity. A perfect spammer lies on the diagonal line on the ROC plot. (b) Contours of equal accuracy (4) and (c) equal spammer score (3). Accuracy This notion of a spammer is quite different for that of the accuracy of an annotator. An annotator with high accuracy is a good annotator but one with low accuracy is not necessarily a spammer. The accuracy is computed as 1 j Accuracyj = Pr[yi = yi ] = j Pr[yi = 1|yi = k]Pr[yi = k] = αj p + β j (1 − p), (4) k=0 where p := Pr[yi = 1] is the prevalence of the positive class. Note that accuracy depends on prevalence. Our proposed spammer score does not depend on prevalence and essentially quantifies the annotator’s inherent discriminatory power. Figure 1(b) shows the contours of equal accuracy on the ROC plot. Note that annotators below the diagonal line (malicious annotators) have low accuracy. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we can detect them and then correct for the flipping. In fact the EM algorithms [3, 7] can correctly flip the labels for the malicious annotators and hence they should not be treated as spammers. Figure 1(c) also shows the contours of equal score for our proposed score and it can be seen that the malicious annotators have a high score and only annotators along the diagonal have a low score (spammers). Log-odds Another interpretation of a spammer can be seen from the log odds. Using Bayes’ rule the posterior log-odds can be written as log j Pr[yi = 1|yi ] Pr[yi = j 0|yi ] = log j Pr[yi |yi = 1] j Pr[yi |yi = 0] + log p . 1−p Pr[y =1|y j ] p If an annotator is a spammer (i.e., (2) holds) then log Pr[yi =0|yi ] = log 1−p . Essentially the annotator j i i provides no information in updating the posterior log-odds and hence does not contribute to the estimation of the actual true label. 3 Spammer score for categorical labels Annotator model Suppose there are K ≥ 2 categories. We introduce a multinomial parameter αj = (αj , . . . , αj ) for each annotator, where c c1 cK K j αj := Pr[yi = k|yi = c] ck αj = 1. ck and k=1 αj ck The term denotes the probability that annotator j assigns class k to an instance given that the true class is c. When K = 2, αj and αj are sensitivity and specificity, respectively. 11 00 Who is a spammer? As earlier a spammer assigns labels randomly, i.e., j j Pr[yi = k|yi ] = Pr[yi = k], ∀k. 3 j j This is equivalent to Pr[yi = k|yi = c] = Pr[yi = k|yi = c ], ∀c, c , k = 1, . . . , K— which means knowing the true class label being c or c does not change the probability of the annotator’s assigned label. This indicates that the annotator j is a spammer if αj = αj k , ∀c, c , k = 1, . . . , K. ck c (5) Let Aj be the K × K confusion rate matrix with entries [Aj ]ck = αck —a spammer would have 0.50 0.50 0.50 all the rows of Aj equal, for example, Aj = 0.25 0.25 0.25 0.25 0.25 0.25 , for a three class categorical annotation problem. Essentially Aj is a rank one matrix of the form Aj = evj , for some column vector vj ∈ RK that satisfies vj e = 1, where e is column vector of ones. In the binary case we had this natural notion of spammer as an annotator for whom αj + β j − 1 was close to zero. One natural way to summarize (5) would be in terms of the distance (Frobenius norm) of the confusion matrix to the closest rank one approximation, i.e, S j := Aj − eˆj v 2 F, (6) where ˆj solves v ˆj = arg min Aj − evj v vj 2 F s.t. vj e = 1. (7) Solving (7) yields ˆj = (1/K)Aj e, which is the mean of the rows of Aj . Then from (6) we have v Sj = I− 1 ee K 2 Aj = F 1 K (αj − αj k )2 . ck c c < . . . < K. Annotator model It is conceptually easier to think of the true label to be binary, that is, yi ∈ {0, 1}. For example in mammography a lesion is either malignant (1) or benign (0) (which can be confirmed by biopsy) and the BIRADS ordinal scale is a means for the radiologist to quantify the uncertainty based on the digital mammogram. The radiologist assigns a higher value of the label if he/she thinks the true label is closer to one. As earlier we characterize each annotator by the sensitivity and the specificity, but the main difference is that we now define the sensitivity and specificity for j each ordinal label (or threshold) k ∈ {1, . . . , K}. Let αj and βk be the sensitivity and specificity k th respectively of the j annotator corresponding to the threshold k, that is, j j j αj = Pr[yi ≥ k | yi = 1] and βk = Pr[yi < k | yi = 0]. k j j Note that αj = 1, β1 = 0 and αj 1 K+1 = 0, βK+1 = 1 from this definition. Hence each annotator j j is parameterized by a set of 2(K − 1) parameters [αj , β2 , . . . , αj , βK ]. This corresponds to an 2 K empirical ROC curve for the annotator (Figure 2). 4 Who is a spammer? As earlier we define an an1 j k=1 notator j to be a spammer if Pr[yi = k|yi = 1] = 0.9 j k=2 0.8 Pr[yi = k|yi = 0] ∀k = 1, . . . , K. Note that from j 0.7 k=3 [ 1−β , α ] the annotation model we have 3 Pr[yi = k | yi = 0.6 j j j 1] = αk − αk+1 and Pr[yi = k | yi = 0] = 0.5 k=4 j j 0.4 βk+1 − βk . This implies that annotator j is a spam0.3 j j mer if αj − αj k k+1 = βk+1 − βk , ∀k = 1, . . . , K, 0.2 j j 0.1 which leads to αj + βk = αj + β1 = 1, ∀k. This 1 k j 0 0 0.2 0.4 0.6 0.8 1 means that for every k, the point (1 − βk , αj ) lies on k 1−Specificity ( β ) the diagonal line in the ROC plot shown in Figure 2. The area under the empirical ROC curve can be comFigure 2: Ordinal labels: An annotator is modK 1 puted as (see Figure 2) AUCj = 2 k=1 (αj + eled by sensitivity/specificity for each threshold. k+1 j j αj )(βk+1 − βk ), and can be used to define the folk lowing spammer score as (2AUCj − 1)2 to rank the different annotators. 3 Sensitivity ( αj ) 3 j 2 K (αj k+1 k=1 j S = + j αj )(βk+1 k − j βk ) −1 (9) With two levels this expression defaults to the binary case. An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 5 Previous work Recently Ipeirotis et.al. [4] proposed a score for categorical labels based on the expected cost of the posterior label. In this section we briefly describe their approach and compare it with our proposed score. For each instance labeled by the annotator they first compute the posterior (soft) label j j Pr[yi = c|yi ] for c = 1, . . . , K, where yi is the label assigned to the ith instance by the j th annotator and yi is the true unknown label. The posterior label is computed via Bayes’ rule as j j j Pr[yi = c|yi ] ∝ Pr[yi |yi = c]Pr[yi = c] = (αj )δ(yi ,k) pc , where pc = Pr[yi = c] is the prevack lence of class c. The score for a spammer is based on the intuition that the posterior label vector j j (Pr[yi = 1|yi ], . . . , Pr[yi = K|yi ]) for a good annotator will have all the probability mass concentrated on single class. For example for a three class problem (with equal prevalence), a posterior label vector of (1, 0, 0) (certain that the class is one) comes from a good annotator while a (1/3, 1/3, 1/3) (complete uncertainty about the class label) comes from spammer. Based on this they define the following score for each annotator 1 Score = N N K K j j costck Pr[yi = k|yi ]Pr[yi = c|yi ] j i=1 . (10) c=1 k=1 where costck is the misclassification cost when an instance of class c is classified as k. Essentially this is capturing some sort of uncertainty of the posterior label averaged over all the instances. Perfect workers have a score Scorej = 0 while spammers will have high score. An entropic version of this score based on similar ideas has also been recently proposed in [5]. Our proposed spammer score differs from this approach in the following aspects: (1) Implicit in the score defined above (10) j is the assumption that an annotator is a spammer when Pr[yi = c|yi ] = Pr[yi = c], i.e., the estimated posterior labels are simply based on the prevalence and do not depend on the observed labels. By j j Bayes’ rule this is equivalent to Pr[yi |yi = c] = Pr[yi ] which is what we have used to define our spammer score. (2) While both notions of a spammer are equivalent, the approach of [4] first computes the posterior labels based on the observed data, the class prevalence and the annotator j j j j This can be seen as follows: Pr[yi = k | yi = 1] = Pr[(yi ≥ k) AND (yi < k + 1) | yi = 1] = Pr[yi ≥ j j j j k | yi = 1] + Pr[yi < k + 1 | yi = 1] − Pr[(yi ≥ k) OR (yi < k + 1) | yi = 1] = Pr[yi ≥ k | yi = j j j 1] − Pr[yi ≥ k + 1 | yi = 1] = αj − αj . Here we used the fact that Pr[(yi ≥ k) OR (yi < k + 1)] = 1. k k+1 3 5 simulated | 500 instances | 30 annotators simulated | 500 instances | 30 annotators 1 12 0.8 Spammer Score 18 0.6 0.5 22 24 23 25 0.3 29 20 0.2 0.4 0.2 30 16 14 0.1 26 21 27 28 19 0 0 13 0 0.2 0.4 0.6 1−Specificity 0.8 1 500 500 500 500 500 500 500 500 500 500 0.4 0.6 500 500 500 1 0.7 500 500 500 500 500 500 500 500 500 500 500 500 500 500 3 1 500 500 500 2 8 510 7 17 4 9 27 8 30 6 3 28 7 10 2 23 22 26 24 5 1 21 29 25 14 12 17 11 18 20 19 15 16 13 4 0.8 Sensitivity 6 9 0.9 15 11 Annotator (a) Simulation setup (b) Annotator ranking Annotator rank (median) via accuracy simulated | 500 instances | 30 annotators Annotator rank (median) via Ipeirotis et.al.[4] simulated | 500 instances | 30 annotators 27 30 30 28 23 22 26 25 24 21 29 25 20 14 17 18 15 20 16 15 11 13 12 19 1 10 5 10 2 7 6 3 5 8 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (c) Comparison with accuracy 30 18 20 16 15 19 13 12 25 11 14 17 25 20 29 21 51 15 26 22 23 24 10 2 7 10 28 6 3 8 30 5 27 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (d) Comparison with Ipeirotis et. al. [4] Figure 3: (a) The simulation setup consisting of 10 good annotators (annotators 1 to 10), 10 spammers (11 to 20), and 10 malicious annotators (21 to 30). (b) The ranking of annotators obtained using the proposed spammer score. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. (c) and (d) Comparison of the median rank obtained via the spammer score with the rank obtained using (c) accuracy and (d) the method proposed by Ipeirotis et. al. [4]. parameters and then computes the expected cost. Our proposed spammer score does not depend on the prevalence of the class. Our score is also directly defined only in terms of the annotator confusion matrix and does not need the observed labels. (3) For the score defined in (10) while perfect annotators have a score of 0 it is not clear what should be a good baseline for a spammer. The authors suggest to compute the baseline by assuming that a worker assigns as label the class with maximum prevalence. Our proposed score has a natural scale with a perfect annotator having a score of 1 and a spammer having a score of 0. (4) However one advantage of the approach in [4] is that they can directly incorporate varied misclassification costs. 6 Experiments Ranking annotators based on the confidence interval As mentioned earlier the annotator model parameters can be estimated using the iterative EM algorithms [3, 7] and these estimated annotator parameters can then be used to compute the spammer score. The spammer score can then be used to rank the annotators. However one commonly observed phenomenon when working with crowdsourced data is that we have a lot of annotators who label only a very few instances. As a result the annotator parameters cannot be reliably estimated for these annotators. In order to factor this uncertainty in the estimation of the model parameters we compute the spammer score for 100 bootstrap replications. Based on this we compute the 95% confidence intervals (CI) for the spammer score for each annotator. We rank the annotators based on the lower limit of the 95% CI. The CIs are wider 6 Table 1: Datasets N is the number of instances. M is the number of annotators. M ∗ is the mean/median number of annotators per instance. N ∗ is the mean/median number of instances labeled by each annotator. Dataset Type N M M∗ N∗ bluebird binary 108 39 39/39 108/108 temp binary 462 76 10/10 61/16 Brief Description wsd categorical/3 177 34 10/10 52/20 sentiment categorical/3 1660 33 6/6 291/175 30 100 10 38 10/10 10/10 bird identification [12] The annotator had to identify whether there was an Indigo Bunting or Blue Grosbeak in the image. event annotation [10] Given a dialogue and a pair of verbs annotators need to label whether the event described by the first verb occurs before or after the second. 30/30 26/20 30 30 30 30 30 30 3 30 30 1 30 20 20 20 20 20 77 117 20 60 Spammer Score 0.4 10 7 9 8 6 5 0 2 0.2 13 31 10 23 29 1 2 4 6 8 9 14 15 17 22 32 5 18 16 19 11 12 20 21 24 25 26 27 28 30 33 34 7 3 0 0.6 20 20 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 20 20 20 20 17 17 40 20 20 100 Spammer Score 0.4 108 108 108 108 108 108 108 108 108 108 108 108 108 0.8 0.6 0.2 17 8 27 30 25 35 1 12 32 37 38 16 22 9 29 15 20 19 5 39 3 21 23 14 2 10 24 7 33 13 36 31 4 34 28 18 11 6 26 0.2 30 77 77 4 108 108 108 108 0.4 0 wosi | 30 instances | 10 annotators 1 0.8 108 108 0.6 108 108 108 Spammer Score 0.8 wsd | 177 instances | 34 annotators 1 80 177 157 177 157 bluebird | 108 instances | 39 annotators 1 word similarity [10] Numeric judgements of word similarity. affect recognition [10] Each annotator is presented with a short headline and asked to rate it on a scale [-100,100] to denote the overall positive or negative valence. 40 40 20 ordinal/[0 10] ordinal[-100 100] 20 20 20 wosi valence word sense disambiguation [10] The labeler is given a paragraph of text containing the word ”president” and asked to label one of the three appropriate senses. irish economic sentiment analysis [1] Articles from three Irish online news sources were annotated by volunteer users as positive, negative, or irrelevant. 20 20 20 20 20 60 20 20 20 40 40 100 0.4 Annotator Annotator 0 1 26 10 18 28 15 5 36 23 12 8 32 31 38 13 17 27 11 2 35 24 19 9 6 30 33 37 14 29 4 3 20 34 22 25 7 16 21 40 20 0.2 26 2 6 11 5 14 3 20 9 22 31 10 12 18 8 13 30 4 1 29 19 17 27 28 21 15 25 23 7 33 16 24 32 10 132 10 360 10 0 13 18 52 75 33 32 12 74 31 51 41 55 7 14 70 42 58 65 43 1 10 47 61 73 25 37 76 67 24 46 54 48 39 56 15 62 68 44 53 64 40 9 28 6 2 57 3 4 5 8 11 16 17 19 20 21 22 23 26 27 29 30 34 35 36 38 45 49 50 59 60 63 66 69 71 72 442 462 452 10 10 0.6 238 171 75 654 20 0.2 0.2 0 12 77 67 374 249 229 453 346 428 0.4 Spammer Score 43 175 119 541 525 437 0.8 917 104 284 0.4 0.6 1211 1099 10 Spammer Score 0.8 572 30 52 402 60 0.6 30 Spammer Score 0.8 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 60 40 20 15 7 7 11 12 35 29 1 87 10 10 10 10 10 10 10 12 1 30 10 10 10 10 10 10 10 10 10 10 10 10 10 valence | 100 instances | 38 annotators 20 22 10 10 10 10 sentiment | 1660 instances | 33 annotators 10 10 30 20 10 Annotator temp | 462 instances | 76 annotators 1 Annotator 10 50 10 10 40 10 70 350 80 40 100 192 190 40 32 60 70 20 20 40 80 20 50 50 50 30 Annotator Annotator Figure 4: Annotator Rankings The rankings obtained for the datasets in Table 1. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. Note that the CIs are wider when the annotator labels only a few instances. when the annotator labels only a few instances. For a crowdsourced labeling task the annotator has to be good and also label a reasonable number of instances in order to be reliably identified. Simulated data We first illustrate our proposed spammer score on simulated binary data (with equal prevalence for both classes) consisting of 500 instances labeled by 30 annotators of varying sensitivity and specificity (see Figure 3(a) for the simulation setup). Of the 30 annotators we have 10 good annotators (annotators 1 to 10 who lie above the diagonal in Figure 3(a)), 10 spammers (annotators 11 to 20 who lie around the diagonal), and 10 malicious annotators (annotators 21 to 30 who lie below the diagonal). Figure 3(b) plots the ranking of annotators obtained using the proposed spammer score with the annotator model parameters estimated via the EM algorithm [3, 7]. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence interval (CI) obtained via bootstrapping are shown. The annotators are ranked based on the lower limit of the 95% CI. As can be seen all the spammers (annotators 11 to 20) have a low spammer score and appear at the bottom of the list. The malicious annotators have higher score than the spammers since we can correct for their flipping. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we detect that they are malicious. Figure 3(c) compares the (median) rank obtained via the spammer score with the (median) rank obtained using accuracy as the score to rank the annotators. While the good annotators are ranked high by both methods the accuracy score gives a low rank to the malicious annotators. Accuracy does not capture the notion of a spammer. Figure 3(d) compares the ranking with the method proposed by Ipeirotis et. al. [4] which gives almost similar rankings as our proposed score. 7 21 23 10 6 35 4 34 1126 18 147 30 3 31 13 2436 33 25 5 2 20 15 39 19 15 20 28 22 299 12 37 16 38 10 32 1 5 27 25 35 30 8 17 0 0 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 40 bluebird | 108 instances | 39 annotators 40 1 6 34 112618 4 31 1013 7 30 2 28 21 5 20 15 39 19 20 15 22 37 16 299 12 38 10 5 8 17 0 0 27 25 35 30 0.6 0.5 35 32 2 0.4 36 11 13 31 24 10 33 28 21 26 18 0 0 40 34 15 19 39 0.1 (a) 22 37 20 38 29 9 0.2 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 6 4 16 0.3 32 1 7 30 25 1 3 14 3 27 5 0.7 24 33 14 36 23 25 12 8 0.9 17 0.8 35 Sensitivity Annotator rank (median) via accuracy bluebird | 108 instances | 39 annotators Annotator rank (median) via Ipeirotis et.al.[4] bluebird | 108 instances | 39 annotators 40 23 0.2 0.4 0.6 1−Specificity (b) 0.8 1 (c) Figure 5: Comparison of the rank obtained via the spammer score with the rank obtained using (a) accuracy and (b) the method proposed by Ipeirotis et. al. [4] for the bluebird binary dataset. (c) The annotator model parameters as estimated by the EM algorithm [3, 7]. 19 25 12 18 7 3 14 20 32 5 8 1 16 20 9 21 15 34 10 31 29 17 28 22 26 2315 5 2 0 0 4 6 13 10 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 30 25 16 19 7 25 8 9 27 14 3 28 17 18 32 5 10 4 2 10 6 1529 31 23 22 21 15 0 0 33 30 11 1 20 5 sentiment | 1660 instances | 33 annotators 24 35 12 20 24 34 26 13 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 33 7 30 15 17 25 28 2719 2223 20 8 1 4 1812 15 13 10 20 32 30 10 3 29 9 31 16 5 6 2 5 14 11 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score 25 21 Annotator rank (median) via Ipeirotis et.al.[4] 25 24 27 Annotator rank (median) via accuracy Annotator rank (median) via accuracy 30 sentiment | 1660 instances | 33 annotators wsd | 177 instances | 34 annotators 33 30 11 Annotator rank (median) via Ipeirotis et.al.[4] wsd | 177 instances | 34 annotators 35 7 30 15 19 17 27 25 21 25 8 12 4 18 20 24 15 20 33 10 3 13 9 28 1 29 23 10 1632 11 14 5 6 2 5 31 30 22 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score Figure 6: Comparison of the median rank obtained via the spammer score with the rank obtained using accuracy and he method proposed by Ipeirotis et. al. [4] for the two categorial datasets in Table 1. Mechanical Turk data We report results on some publicly available linguistic and image annotation data collected using the Amazon’s Mechanical Turk (AMT) and other sources. Table 1 summarizes the datasets. Figure 4 plots the spammer scores and rankings obtained. The mean and the 95% CI obtained via bootstrapping are also shown. The number at the top of the CI bar shows the number of instances annotated by that annotator. The rankings are based on the lower limit of the 95% CI which factors the number of instances labeled by the annotator into the ranking. An annotator who labels only a few instances will have very wide CI. Some annotators who label only a few instances may have a high mean spammer score but the CI will be wide and hence ranked lower. Ideally we would like to have annotators with a high score and at the same time label a lot of instances so that we can reliablly identify them. The authors [1] for the sentiment dataset shared with us some of the qualitative observations regarding the annotators and they somewhat agree with our rankings. For example the authors made the following comments about Annotator 7 ”Quirky annotator - had a lot of debate about what was the meaning of the annotation question. I’d say he changed his labeling strategy at least once during the process”. Our proposed score gave a low rank to this annotator. Comparison with other approaches Figure 5 and 6 compares the proposed ranking with the rank obtained using accuracy and the method proposed by Ipeirotis et. al. [4] for some binary and categorical datasets in Table 1. Our proposed ranking is somewhat similar to that obtained by Ipeirotis et. al. [4] but accuracy does not quite capture the notion of spammer. For example for the bluebird dataset for annotator 21 (see Figure 5(a)) accuracy ranks it at the bottom of the list while the proposed score puts is in the middle of the list. From the estimated model parameters it can be seen that annotator 21 actually flips the labels (below the diagonal in Figure 5(c)) but is a good annotator. 7 Conclusions We proposed a score to rank annotators for crowdsourced binary, categorical, and ordinal labeling tasks. The obtained rankings and the scores can be used to allocate monetary bonuses to be paid to different annotators and also to eliminate spammers from further labeling tasks. A mechanism to rank annotators should be desirable feature of any crowdsourcing service. The proposed score should also be useful to specify the prior for Bayesian approaches to consolidate annotations. 8 References [1] A. Brew, D. Greene, and P. Cunningham. Using crowdsourcing and active learning to track sentiment in online media. In Proceedings of the 6th Conference on Prestigious Applications of Intelligent Systems (PAIS’10), 2010. [2] B. Carpenter. Multilevel bayesian models of categorical data annotation. Technical Report available at http://lingpipe-blog.com/lingpipe-white-papers/, 2008. [3] A. P. Dawid and A. M. Skene. Maximum likeihood estimation of observer error-rates using the EM algorithm. Applied Statistics, 28(1):20–28, 1979. [4] P. G. Ipeirotis, F. Provost, and J. Wang. Quality management on Amazon Mechanical Turk. In Proceedings of the ACM SIGKDD Workshop on Human Computation (HCOMP’10), pages 64–67, 2010. [5] V. C. Raykar and S. Yu. An entropic score to rank annotators for crowdsourced labelling tasks. In Proceedings of the Third National Conference on Computer Vision, Pattern Recognition, Image Processing and Graphics (NCVPRIPG), 2011. [6] V. C. Raykar, S. Yu, L .H. Zhao, A. Jerebko, C. Florin, G. H. Valadez, L. Bogoni, and L. Moy. Supervised learning from multiple experts: Whom to trust when everyone lies a bit. In Proceedings of the 26th International Conference on Machine Learning (ICML 2009), pages 889– 896, 2009. [7] V. C. Raykar, S. Yu, L. H. Zhao, G. H. Valadez, C. Florin, L. Bogoni, and L. Moy. Learning from crowds. Journal of Machine Learning Research, 11:1297–1322, April 2010. [8] V. S. Sheng, F. Provost, and P. G. Ipeirotis. Get another label? Improving data quality and data mining using multiple, noisy labelers. In Proceedings of the 14th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pages 614–622, 2008. [9] P. Smyth, U. Fayyad, M. Burl, P. Perona, and P. Baldi. Inferring ground truth from subjective labelling of venus images. In Advances in Neural Information Processing Systems 7, pages 1085–1092. 1995. [10] R. Snow, B. O’Connor, D. Jurafsky, and A. Y. Ng. Cheap and Fast—but is it good? Evaluating Non-Expert Annotations for Natural Language Tasks. In Proceedings of the Conference on Empirical Methods in Natural Language Processing (EMNLP ’08), pages 254–263, 2008. [11] A. Sorokin and D. Forsyth. Utility data annotation with Amazon Mechanical Turk. In Proceedings of the First IEEE Workshop on Internet Vision at CVPR 08, pages 1–8, 2008. [12] P. Welinder, S. Branson, S. Belongie, and P. Perona. The multidimensional wisdom of crowds. In Advances in Neural Information Processing Systems 23, pages 2424–2432. 2010. [13] J. Whitehill, P. Ruvolo, T. Wu, J. Bergsma, and J. Movellan. Whose vote should count more: Optimal integration of labels from labelers of unknown expertise. In Advances in Neural Information Processing Systems 22, pages 2035–2043. 2009. [14] Y. Yan, R. Rosales, G. Fung, M. Schmidt, G. Hermosillo, L. Bogoni, L. Moy, and J. Dy. Modeling annotator expertise: Learning when everybody knows a bit of something. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics (AISTATS 2010), pages 932–939, 2010. 9
Reference: text
sentIndex sentText sentNum sentScore
1 In this paper we formalize the notion of a spammer and define a score which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a high score close to one. [sent-12, score-1.614]
2 In this paper we are interested in ranking annotators based on how spammer like each annotator is. [sent-19, score-1.61]
3 In our context a spammer is a low quality annotator who assigns random labels (maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator). [sent-20, score-1.966]
4 More specifically we define a scalar metric which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a score close to one (see Figure 4). [sent-25, score-0.953]
5 We summarize the multiple parameters corresponding to each annotator into a single score indicative of how spammer like the annotator is. [sent-26, score-1.882]
6 While this spammer score was implicit for binary labels in earlier works [3, 9, 2, 6] the extension to categorical and ordinal labels is novel and is quite different from the accuracy computed from the confusion rate matrix. [sent-27, score-0.916]
7 While we obtain somewhat similar annotator rankings, we differ from this work in that our score is directly defined in terms of the annotator parameters (see § 5 for more details). [sent-29, score-1.388]
8 We first specify the annotator model used, formalize the notion of a spammer, and propose an appropriate score in terms of the annotator model parameters. [sent-32, score-1.411]
9 We do not dwell too much on the estimation of the annotator model parameters. [sent-33, score-0.622]
10 These parameters can either be estimated directly using known gold standard 1 or the iterative algorithms that estimate the annotator model parameters without actually knowing the gold standard [3, 9, 2, 6, 7]. [sent-34, score-0.682]
11 In the experimental section (§ 6) we obtain rankings for the annotators using the proposed spammer scores on some publicly available data from different domains. [sent-35, score-1.005]
12 2 Spammer score for crowdsourced binary labels j Annotator model Let yi ∈ {0, 1} be the label assigned to the ith instance by the j th annotator, and let yi ∈ {0, 1} be the actual (unobserved) binary label. [sent-36, score-0.617]
13 We model the accuracy of the annotator separately on the positive and the negative examples. [sent-37, score-0.654]
14 If the true label is one, the sensitivity (true positive rate) αj for the j th annotator is defined as the probability that the annotator labels it as one. [sent-38, score-1.384]
15 On the other hand, if the true label is zero, the specificity (1−false positive rate) β j is defined as the probability that annotator labels it as zero. [sent-40, score-0.718]
16 Extensions of this basic model have been proposed to include item level difficulty [2, 13] and also to model the annotator performance based on the feature vector [14]. [sent-42, score-0.634]
17 Based on many instances labeled by multiple annotators the maximum likelihood estimator for the annotator parameters (αj , β j ) and also the consensus ground truth (yi ) can be estimated iteratively [3, 7] via the Expectation Maximization (EM) algorithm. [sent-44, score-1.203]
18 Intuitively, a spammer assigns labels randomly—maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator. [sent-47, score-1.331]
19 More precisely an annotator is a spammer if the probability j of observed label yi being one given the true label yi is independent of the true label, i. [sent-48, score-1.506]
20 This means that the annotator is assigning labels randomly by flipping a coin with bias without actually looking at the data. [sent-51, score-0.678]
21 (1) j Pr[yi = 1] (2) Hence in the context of the annotator model defined earlier a perfect spammer is an annotator for whom αj + β j − 1 = 0. [sent-53, score-1.769]
22 If αj + β j − 1 < 0 then the annotators lies below the diagonal line and is a malicious annotator who flips the labels. [sent-55, score-1.172]
23 Note that a malicious annotator has discriminatory power if we can detect them and flip their labels. [sent-56, score-0.716]
24 Hence we define the spammer score for an annotator as S j = (αj + β j − 1)2 (3) An annotator is a spammer if S j is close to zero. [sent-58, score-2.376]
25 Good annotators have S j > 0 while a perfect annotator has S j = 1. [sent-59, score-1.107]
26 8 1 (c) Spammer score Figure 1: (a) For binary labels an annotator is modeled by his/her sensitivity and specificity. [sent-155, score-0.88]
27 (b) Contours of equal accuracy (4) and (c) equal spammer score (3). [sent-157, score-0.696]
28 An annotator with high accuracy is a good annotator but one with low accuracy is not necessarily a spammer. [sent-159, score-1.321]
29 Our proposed spammer score does not depend on prevalence and essentially quantifies the annotator’s inherent discriminatory power. [sent-162, score-0.713]
30 The malicious annotators are good annotators but they flip their labels and as such are not spammers if we can detect them and then correct for the flipping. [sent-165, score-1.197]
31 In fact the EM algorithms [3, 7] can correctly flip the labels for the malicious annotators and hence they should not be treated as spammers. [sent-166, score-0.59]
32 Figure 1(c) also shows the contours of equal score for our proposed score and it can be seen that the malicious annotators have a high score and only annotators along the diagonal have a low score (spammers). [sent-167, score-1.64]
33 1−p Pr[y =1|y j ] p If an annotator is a spammer (i. [sent-170, score-1.116]
34 Essentially the annotator j i i provides no information in updating the posterior log-odds and hence does not contribute to the estimation of the actual true label. [sent-173, score-0.639]
35 ck and k=1 αj ck The term denotes the probability that annotator j assigns class k to an instance given that the true class is c. [sent-179, score-0.693]
36 This indicates that the annotator j is a spammer if αj = αj k , ∀c, c , k = 1, . [sent-189, score-1.116]
37 In the binary case we had this natural notion of spammer as an annotator for whom αj + β j − 1 was close to zero. [sent-204, score-1.143]
38 As earlier we characterize each annotator by the sensitivity and the specificity, but the main difference is that we now define the sensitivity and specificity for j each ordinal label (or threshold) k ∈ {1, . [sent-218, score-0.805]
39 Let αj and βk be the sensitivity and specificity k th respectively of the j annotator corresponding to the threshold k, that is, j j j αj = Pr[yi ≥ k | yi = 1] and βk = Pr[yi < k | yi = 0]. [sent-222, score-0.976]
40 Hence each annotator j j is parameterized by a set of 2(K − 1) parameters [αj , β2 , . [sent-224, score-0.622]
41 This corresponds to an 2 K empirical ROC curve for the annotator (Figure 2). [sent-228, score-0.622]
42 The area under the empirical ROC curve can be comFigure 2: Ordinal labels: An annotator is modK 1 puted as (see Figure 2) AUCj = 2 k=1 (αj + eled by sensitivity/specificity for each threshold. [sent-253, score-0.622]
43 k+1 j j αj )(βk+1 − βk ), and can be used to define the folk lowing spammer score as (2AUCj − 1)2 to rank the different annotators. [sent-254, score-0.707]
44 An annotator is a spammer if S j is close to zero. [sent-256, score-1.116]
45 Good annotators have S j > 0 while a perfect annotator has S j = 1. [sent-257, score-1.107]
46 For each instance labeled by the annotator they first compute the posterior (soft) label j j Pr[yi = c|yi ] for c = 1, . [sent-262, score-0.693]
47 , K, where yi is the label assigned to the ith instance by the j th annotator and yi is the true unknown label. [sent-265, score-0.972]
48 The score for a spammer is based on the intuition that the posterior label vector j j (Pr[yi = 1|yi ], . [sent-267, score-0.695]
49 , Pr[yi = K|yi ]) for a good annotator will have all the probability mass concentrated on single class. [sent-270, score-0.635]
50 For example for a three class problem (with equal prevalence), a posterior label vector of (1, 0, 0) (certain that the class is one) comes from a good annotator while a (1/3, 1/3, 1/3) (complete uncertainty about the class label) comes from spammer. [sent-271, score-0.705]
51 Based on this they define the following score for each annotator 1 Score = N N K K j j costck Pr[yi = k|yi ]Pr[yi = c|yi ] j i=1 . [sent-272, score-0.781]
52 Our proposed spammer score differs from this approach in the following aspects: (1) Implicit in the score defined above (10) j is the assumption that an annotator is a spammer when Pr[yi = c|yi ] = Pr[yi = c], i. [sent-277, score-1.91]
53 k k+1 3 5 simulated | 500 instances | 30 annotators simulated | 500 instances | 30 annotators 1 12 0. [sent-283, score-1.08]
54 9 15 11 Annotator (a) Simulation setup (b) Annotator ranking Annotator rank (median) via accuracy simulated | 500 instances | 30 annotators Annotator rank (median) via Ipeirotis et. [sent-300, score-0.764]
55 [4] Figure 3: (a) The simulation setup consisting of 10 good annotators (annotators 1 to 10), 10 spammers (11 to 20), and 10 malicious annotators (21 to 30). [sent-304, score-1.13]
56 (b) The ranking of annotators obtained using the proposed spammer score. [sent-305, score-1.0]
57 The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. [sent-306, score-0.67]
58 The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. [sent-307, score-0.654]
59 (c) and (d) Comparison of the median rank obtained via the spammer score with the rank obtained using (c) accuracy and (d) the method proposed by Ipeirotis et. [sent-310, score-0.895]
60 Our proposed spammer score does not depend on the prevalence of the class. [sent-314, score-0.698]
61 Our score is also directly defined only in terms of the annotator confusion matrix and does not need the observed labels. [sent-315, score-0.793]
62 (3) For the score defined in (10) while perfect annotators have a score of 0 it is not clear what should be a good baseline for a spammer. [sent-316, score-0.786]
63 Our proposed score has a natural scale with a perfect annotator having a score of 1 and a spammer having a score of 0. [sent-318, score-1.579]
64 6 Experiments Ranking annotators based on the confidence interval As mentioned earlier the annotator model parameters can be estimated using the iterative EM algorithms [3, 7] and these estimated annotator parameters can then be used to compute the spammer score. [sent-320, score-2.236]
65 The spammer score can then be used to rank the annotators. [sent-321, score-0.707]
66 As a result the annotator parameters cannot be reliably estimated for these annotators. [sent-323, score-0.632]
67 In order to factor this uncertainty in the estimation of the model parameters we compute the spammer score for 100 bootstrap replications. [sent-324, score-0.654]
68 Based on this we compute the 95% confidence intervals (CI) for the spammer score for each annotator. [sent-325, score-0.638]
69 8 wsd | 177 instances | 34 annotators 1 80 177 157 177 157 bluebird | 108 instances | 39 annotators 1 word similarity [10] Numeric judgements of word similarity. [sent-345, score-1.152]
70 affect recognition [10] Each annotator is presented with a short headline and asked to rate it on a scale [-100,100] to denote the overall positive or negative valence. [sent-346, score-0.622]
71 The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. [sent-362, score-0.67]
72 The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. [sent-363, score-0.654]
73 Note that the CIs are wider when the annotator labels only a few instances. [sent-366, score-0.678]
74 For a crowdsourced labeling task the annotator has to be good and also label a reasonable number of instances in order to be reliably identified. [sent-368, score-0.802]
75 Simulated data We first illustrate our proposed spammer score on simulated binary data (with equal prevalence for both classes) consisting of 500 instances labeled by 30 annotators of varying sensitivity and specificity (see Figure 3(a) for the simulation setup). [sent-369, score-1.323]
76 Of the 30 annotators we have 10 good annotators (annotators 1 to 10 who lie above the diagonal in Figure 3(a)), 10 spammers (annotators 11 to 20 who lie around the diagonal), and 10 malicious annotators (annotators 21 to 30 who lie below the diagonal). [sent-370, score-1.612]
77 Figure 3(b) plots the ranking of annotators obtained using the proposed spammer score with the annotator model parameters estimated via the EM algorithm [3, 7]. [sent-371, score-1.789]
78 The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. [sent-372, score-0.67]
79 The mean spammer score and the 95% confidence interval (CI) obtained via bootstrapping are shown. [sent-373, score-0.651]
80 As can be seen all the spammers (annotators 11 to 20) have a low spammer score and appear at the bottom of the list. [sent-375, score-0.755]
81 The malicious annotators have higher score than the spammers since we can correct for their flipping. [sent-376, score-0.795]
82 The malicious annotators are good annotators but they flip their labels and as such are not spammers if we detect that they are malicious. [sent-377, score-1.197]
83 Figure 3(c) compares the (median) rank obtained via the spammer score with the (median) rank obtained using accuracy as the score to rank the annotators. [sent-378, score-1.034]
84 While the good annotators are ranked high by both methods the accuracy score gives a low rank to the malicious annotators. [sent-379, score-0.81]
85 2 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 6 4 16 0. [sent-389, score-0.72]
86 8 35 Sensitivity Annotator rank (median) via accuracy bluebird | 108 instances | 39 annotators Annotator rank (median) via Ipeirotis et. [sent-393, score-0.774]
87 8 1 (c) Figure 5: Comparison of the rank obtained via the spammer score with the rank obtained using (a) accuracy and (b) the method proposed by Ipeirotis et. [sent-399, score-0.833]
88 (c) The annotator model parameters as estimated by the EM algorithm [3, 7]. [sent-402, score-0.632]
89 [4] 25 24 27 Annotator rank (median) via accuracy Annotator rank (median) via accuracy 30 sentiment | 1660 instances | 33 annotators wsd | 177 instances | 34 annotators 33 30 11 Annotator rank (median) via Ipeirotis et. [sent-405, score-1.425]
90 The rankings are based on the lower limit of the 95% CI which factors the number of instances labeled by the annotator into the ranking. [sent-415, score-0.729]
91 An annotator who labels only a few instances will have very wide CI. [sent-416, score-0.738]
92 Some annotators who label only a few instances may have a high mean spammer score but the CI will be wide and hence ranked lower. [sent-417, score-1.222]
93 Ideally we would like to have annotators with a high score and at the same time label a lot of instances so that we can reliablly identify them. [sent-418, score-0.721]
94 For example the authors made the following comments about Annotator 7 ”Quirky annotator - had a lot of debate about what was the meaning of the annotation question. [sent-420, score-0.655]
95 For example for the bluebird dataset for annotator 21 (see Figure 5(a)) accuracy ranks it at the bottom of the list while the proposed score puts is in the middle of the list. [sent-429, score-0.862]
96 From the estimated model parameters it can be seen that annotator 21 actually flips the labels (below the diagonal in Figure 5(c)) but is a good annotator. [sent-430, score-0.717]
97 7 Conclusions We proposed a score to rank annotators for crowdsourced binary, categorical, and ordinal labeling tasks. [sent-431, score-0.801]
98 The obtained rankings and the scores can be used to allocate monetary bonuses to be paid to different annotators and also to eliminate spammers from further labeling tasks. [sent-432, score-0.669]
99 An entropic score to rank annotators for crowdsourced labelling tasks. [sent-464, score-0.729]
100 Modeling annotator expertise: Learning when everybody knows a bit of something. [sent-551, score-0.622]
wordName wordTfidf (topN-words)
[('annotator', 0.622), ('spammer', 0.494), ('annotators', 0.466), ('yi', 0.155), ('score', 0.144), ('pr', 0.132), ('spammers', 0.117), ('ipeirotis', 0.078), ('rank', 0.069), ('malicious', 0.068), ('median', 0.062), ('instances', 0.06), ('labels', 0.056), ('bluebird', 0.052), ('prevalence', 0.048), ('sensitivity', 0.044), ('ordinal', 0.043), ('aj', 0.041), ('label', 0.04), ('crowdsourced', 0.039), ('specificity', 0.039), ('categorical', 0.038), ('sentiment', 0.037), ('rankings', 0.033), ('accuracy', 0.032), ('crowdsourcing', 0.029), ('ranking', 0.028), ('labeling', 0.028), ('roc', 0.027), ('city', 0.027), ('confusion', 0.027), ('wsd', 0.026), ('ck', 0.026), ('gold', 0.025), ('ci', 0.025), ('maybe', 0.024), ('contours', 0.023), ('raykar', 0.022), ('spammy', 0.022), ('annotation', 0.022), ('perfect', 0.019), ('ip', 0.019), ('assigns', 0.019), ('mechanical', 0.019), ('ranked', 0.018), ('bogoni', 0.018), ('consensus', 0.018), ('annotations', 0.018), ('posterior', 0.017), ('amazon', 0.017), ('em', 0.016), ('diagonal', 0.016), ('bootstrap', 0.016), ('services', 0.016), ('annotated', 0.015), ('bonuses', 0.015), ('costck', 0.015), ('discriminatory', 0.015), ('evj', 0.015), ('malvern', 0.015), ('pretending', 0.015), ('radiologist', 0.015), ('siemens', 0.015), ('temp', 0.015), ('valence', 0.015), ('wosi', 0.015), ('binary', 0.014), ('simulated', 0.014), ('labeled', 0.014), ('notion', 0.013), ('vj', 0.013), ('good', 0.013), ('equal', 0.013), ('quality', 0.013), ('healthcare', 0.013), ('bcr', 0.013), ('bot', 0.013), ('bar', 0.013), ('turk', 0.013), ('via', 0.013), ('proposed', 0.012), ('earlier', 0.012), ('florin', 0.012), ('provost', 0.012), ('valadez', 0.012), ('irish', 0.012), ('advent', 0.012), ('lot', 0.011), ('entropic', 0.011), ('cis', 0.011), ('word', 0.011), ('detect', 0.011), ('cheap', 0.011), ('ips', 0.011), ('estimated', 0.01), ('formalize', 0.01), ('dence', 0.01), ('ranges', 0.01), ('monetary', 0.01), ('workers', 0.01)]
simIndex simValue paperId paperTitle
same-paper 1 1.0000002 232 nips-2011-Ranking annotators for crowdsourced labeling tasks
Author: Vikas C. Raykar, Shipeng Yu
Abstract: With the advent of crowdsourcing services it has become quite cheap and reasonably effective to get a dataset labeled by multiple annotators in a short amount of time. Various methods have been proposed to estimate the consensus labels by correcting for the bias of annotators with different kinds of expertise. Often we have low quality annotators or spammers–annotators who assign labels randomly (e.g., without actually looking at the instance). Spammers can make the cost of acquiring labels very expensive and can potentially degrade the quality of the consensus labels. In this paper we formalize the notion of a spammer and define a score which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a high score close to one. 1 Spammers in crowdsourced labeling tasks Annotating an unlabeled dataset is one of the bottlenecks in using supervised learning to build good predictive models. Getting a dataset labeled by experts can be expensive and time consuming. With the advent of crowdsourcing services (Amazon’s Mechanical Turk being a prime example) it has become quite easy and inexpensive to acquire labels from a large number of annotators in a short amount of time (see [8], [10], and [11] for some computer vision and natural language processing case studies). One drawback of most crowdsourcing services is that we do not have tight control over the quality of the annotators. The annotators can come from a diverse pool including genuine experts, novices, biased annotators, malicious annotators, and spammers. Hence in order to get good quality labels requestors typically get each instance labeled by multiple annotators and these multiple annotations are then consolidated either using a simple majority voting or more sophisticated methods that model and correct for the annotator biases [3, 9, 6, 7, 14] and/or task complexity [2, 13, 12]. In this paper we are interested in ranking annotators based on how spammer like each annotator is. In our context a spammer is a low quality annotator who assigns random labels (maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator). Spammers can significantly increase the cost of acquiring annotations (since they need to be paid) and at the same time decrease the accuracy of the final consensus labels. A mechanism to detect and eliminate spammers is a desirable feature for any crowdsourcing market place. For example one can give monetary bonuses to good annotators and deny payments to spammers. The main contribution of this paper is to formalize the notion of a spammer for binary, categorical, and ordinal labeling tasks. More specifically we define a scalar metric which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a score close to one (see Figure 4). We summarize the multiple parameters corresponding to each annotator into a single score indicative of how spammer like the annotator is. While this spammer score was implicit for binary labels in earlier works [3, 9, 2, 6] the extension to categorical and ordinal labels is novel and is quite different from the accuracy computed from the confusion rate matrix. An attempt to quantify the quality of the workers based on the confusion matrix was recently made by [4] where they transformed the observed labels into posterior soft labels based on the estimated confusion 1 matrix. While we obtain somewhat similar annotator rankings, we differ from this work in that our score is directly defined in terms of the annotator parameters (see § 5 for more details). The rest of the paper is organized as follows. For ease of exposition we start with binary labels (§ 2) and later extend it to categorical (§ 3) and ordinal labels (§ 4). We first specify the annotator model used, formalize the notion of a spammer, and propose an appropriate score in terms of the annotator model parameters. We do not dwell too much on the estimation of the annotator model parameters. These parameters can either be estimated directly using known gold standard 1 or the iterative algorithms that estimate the annotator model parameters without actually knowing the gold standard [3, 9, 2, 6, 7]. In the experimental section (§ 6) we obtain rankings for the annotators using the proposed spammer scores on some publicly available data from different domains. 2 Spammer score for crowdsourced binary labels j Annotator model Let yi ∈ {0, 1} be the label assigned to the ith instance by the j th annotator, and let yi ∈ {0, 1} be the actual (unobserved) binary label. We model the accuracy of the annotator separately on the positive and the negative examples. If the true label is one, the sensitivity (true positive rate) αj for the j th annotator is defined as the probability that the annotator labels it as one. j αj := Pr[yi = 1|yi = 1]. On the other hand, if the true label is zero, the specificity (1−false positive rate) β j is defined as the probability that annotator labels it as zero. j β j := Pr[yi = 0|yi = 0]. Extensions of this basic model have been proposed to include item level difficulty [2, 13] and also to model the annotator performance based on the feature vector [14]. For simplicity we use the basic model proposed in [7] in our formulation. Based on many instances labeled by multiple annotators the maximum likelihood estimator for the annotator parameters (αj , β j ) and also the consensus ground truth (yi ) can be estimated iteratively [3, 7] via the Expectation Maximization (EM) algorithm. The EM algorithm iteratively establishes a particular gold standard (initialized via majority voting), measures the performance of the annotators given that gold standard (M-step), and refines the gold standard based on the performance measures (E-step). Who is a spammer? Intuitively, a spammer assigns labels randomly—maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator. More precisely an annotator is a spammer if the probability j of observed label yi being one given the true label yi is independent of the true label, i.e., j j Pr[yi = 1|yi ] = Pr[yi = 1]. This means that the annotator is assigning labels randomly by flipping a coin with bias without actually looking at the data. Equivalently (1) can be written as j j Pr[yi = 1|yi = 1] = Pr[yi = 1|yi = 0] which implies αj = 1 − β j . (1) j Pr[yi = 1] (2) Hence in the context of the annotator model defined earlier a perfect spammer is an annotator for whom αj + β j − 1 = 0. This corresponds to the diagonal line on the Receiver Operating Characteristic (ROC) plot (see Figure 1(a)) 2 . If αj + β j − 1 < 0 then the annotators lies below the diagonal line and is a malicious annotator who flips the labels. Note that a malicious annotator has discriminatory power if we can detect them and flip their labels. In fact the methods proposed in [3, 7] can automatically flip the labels for the malicious annotators. Hence we define the spammer score for an annotator as S j = (αj + β j − 1)2 (3) An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 1 One of the commonly used strategy to filter out spammers is to inject some items into the annotations with known labels. This is the strategy used by CrowdFlower (http://crowdflower.com/docs/gold). 2 Also note that (αj + β j )/2 is equal to the area shown in the plot and can be considered as a non-parametric approximation to the area under the ROC curve (AUC) based on one observed point. It is also equal to the Balanced Classification Rate (BCR). So a spammer can also be defined as having BCR or AUC equal to 0.5. 2 Equal accuracy contours (prevalence=0.5) 0.9 1 Good Annotators Biased Annotators 0.8 0.9 Sensitivity Sensitivity ( αj ) Spammers 0.5 0.4 j 0.3 j 0.6 0.5 0.4 4 0. 5 0. 3 0. 7 0. 6 0. 4 0. 5 0. 2 0. 3 0. Malicious Annotators 0.2 0.4 0.6 1−Specificity ( βj ) 0.8 0.1 1 0. 0. 2 0. 3 0. 1 0. 0.6 0.5 1 0. 2 0. 2 0. 1 0. 0.4 3 1 0. 0. 2 0. 4 0. 0.2 4 0. 5 0. 0 0 1 3 0. 0.3 0.2 Biased Annotators 4 0.7 6 0. 4 0. 0.8 7 0.3 Area = (α +β )/2 0.2 0 0 0.9 0. 8 0. 8 0. 0.7 6 0. .5 0 5 0. 0.7 [ 1−βj, αj ] 0.6 0.1 6 0. 0.8 0.7 Equal spammer score contours 1 7 0. 8 0. 9 0. Sensitivity 1 (a) Binary annotator model 0.1 1 0. 2 0. 3 0. 0.2 0.4 0.6 1−Specificity 0.8 1 1 0. 0 0 (b) Accuracy 0.2 3 0. 4 0. 0.4 0.6 1−Specificity 5 0. .6 7 0 0. 8 0. 0.8 1 (c) Spammer score Figure 1: (a) For binary labels an annotator is modeled by his/her sensitivity and specificity. A perfect spammer lies on the diagonal line on the ROC plot. (b) Contours of equal accuracy (4) and (c) equal spammer score (3). Accuracy This notion of a spammer is quite different for that of the accuracy of an annotator. An annotator with high accuracy is a good annotator but one with low accuracy is not necessarily a spammer. The accuracy is computed as 1 j Accuracyj = Pr[yi = yi ] = j Pr[yi = 1|yi = k]Pr[yi = k] = αj p + β j (1 − p), (4) k=0 where p := Pr[yi = 1] is the prevalence of the positive class. Note that accuracy depends on prevalence. Our proposed spammer score does not depend on prevalence and essentially quantifies the annotator’s inherent discriminatory power. Figure 1(b) shows the contours of equal accuracy on the ROC plot. Note that annotators below the diagonal line (malicious annotators) have low accuracy. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we can detect them and then correct for the flipping. In fact the EM algorithms [3, 7] can correctly flip the labels for the malicious annotators and hence they should not be treated as spammers. Figure 1(c) also shows the contours of equal score for our proposed score and it can be seen that the malicious annotators have a high score and only annotators along the diagonal have a low score (spammers). Log-odds Another interpretation of a spammer can be seen from the log odds. Using Bayes’ rule the posterior log-odds can be written as log j Pr[yi = 1|yi ] Pr[yi = j 0|yi ] = log j Pr[yi |yi = 1] j Pr[yi |yi = 0] + log p . 1−p Pr[y =1|y j ] p If an annotator is a spammer (i.e., (2) holds) then log Pr[yi =0|yi ] = log 1−p . Essentially the annotator j i i provides no information in updating the posterior log-odds and hence does not contribute to the estimation of the actual true label. 3 Spammer score for categorical labels Annotator model Suppose there are K ≥ 2 categories. We introduce a multinomial parameter αj = (αj , . . . , αj ) for each annotator, where c c1 cK K j αj := Pr[yi = k|yi = c] ck αj = 1. ck and k=1 αj ck The term denotes the probability that annotator j assigns class k to an instance given that the true class is c. When K = 2, αj and αj are sensitivity and specificity, respectively. 11 00 Who is a spammer? As earlier a spammer assigns labels randomly, i.e., j j Pr[yi = k|yi ] = Pr[yi = k], ∀k. 3 j j This is equivalent to Pr[yi = k|yi = c] = Pr[yi = k|yi = c ], ∀c, c , k = 1, . . . , K— which means knowing the true class label being c or c does not change the probability of the annotator’s assigned label. This indicates that the annotator j is a spammer if αj = αj k , ∀c, c , k = 1, . . . , K. ck c (5) Let Aj be the K × K confusion rate matrix with entries [Aj ]ck = αck —a spammer would have 0.50 0.50 0.50 all the rows of Aj equal, for example, Aj = 0.25 0.25 0.25 0.25 0.25 0.25 , for a three class categorical annotation problem. Essentially Aj is a rank one matrix of the form Aj = evj , for some column vector vj ∈ RK that satisfies vj e = 1, where e is column vector of ones. In the binary case we had this natural notion of spammer as an annotator for whom αj + β j − 1 was close to zero. One natural way to summarize (5) would be in terms of the distance (Frobenius norm) of the confusion matrix to the closest rank one approximation, i.e, S j := Aj − eˆj v 2 F, (6) where ˆj solves v ˆj = arg min Aj − evj v vj 2 F s.t. vj e = 1. (7) Solving (7) yields ˆj = (1/K)Aj e, which is the mean of the rows of Aj . Then from (6) we have v Sj = I− 1 ee K 2 Aj = F 1 K (αj − αj k )2 . ck c c < . . . < K. Annotator model It is conceptually easier to think of the true label to be binary, that is, yi ∈ {0, 1}. For example in mammography a lesion is either malignant (1) or benign (0) (which can be confirmed by biopsy) and the BIRADS ordinal scale is a means for the radiologist to quantify the uncertainty based on the digital mammogram. The radiologist assigns a higher value of the label if he/she thinks the true label is closer to one. As earlier we characterize each annotator by the sensitivity and the specificity, but the main difference is that we now define the sensitivity and specificity for j each ordinal label (or threshold) k ∈ {1, . . . , K}. Let αj and βk be the sensitivity and specificity k th respectively of the j annotator corresponding to the threshold k, that is, j j j αj = Pr[yi ≥ k | yi = 1] and βk = Pr[yi < k | yi = 0]. k j j Note that αj = 1, β1 = 0 and αj 1 K+1 = 0, βK+1 = 1 from this definition. Hence each annotator j j is parameterized by a set of 2(K − 1) parameters [αj , β2 , . . . , αj , βK ]. This corresponds to an 2 K empirical ROC curve for the annotator (Figure 2). 4 Who is a spammer? As earlier we define an an1 j k=1 notator j to be a spammer if Pr[yi = k|yi = 1] = 0.9 j k=2 0.8 Pr[yi = k|yi = 0] ∀k = 1, . . . , K. Note that from j 0.7 k=3 [ 1−β , α ] the annotation model we have 3 Pr[yi = k | yi = 0.6 j j j 1] = αk − αk+1 and Pr[yi = k | yi = 0] = 0.5 k=4 j j 0.4 βk+1 − βk . This implies that annotator j is a spam0.3 j j mer if αj − αj k k+1 = βk+1 − βk , ∀k = 1, . . . , K, 0.2 j j 0.1 which leads to αj + βk = αj + β1 = 1, ∀k. This 1 k j 0 0 0.2 0.4 0.6 0.8 1 means that for every k, the point (1 − βk , αj ) lies on k 1−Specificity ( β ) the diagonal line in the ROC plot shown in Figure 2. The area under the empirical ROC curve can be comFigure 2: Ordinal labels: An annotator is modK 1 puted as (see Figure 2) AUCj = 2 k=1 (αj + eled by sensitivity/specificity for each threshold. k+1 j j αj )(βk+1 − βk ), and can be used to define the folk lowing spammer score as (2AUCj − 1)2 to rank the different annotators. 3 Sensitivity ( αj ) 3 j 2 K (αj k+1 k=1 j S = + j αj )(βk+1 k − j βk ) −1 (9) With two levels this expression defaults to the binary case. An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 5 Previous work Recently Ipeirotis et.al. [4] proposed a score for categorical labels based on the expected cost of the posterior label. In this section we briefly describe their approach and compare it with our proposed score. For each instance labeled by the annotator they first compute the posterior (soft) label j j Pr[yi = c|yi ] for c = 1, . . . , K, where yi is the label assigned to the ith instance by the j th annotator and yi is the true unknown label. The posterior label is computed via Bayes’ rule as j j j Pr[yi = c|yi ] ∝ Pr[yi |yi = c]Pr[yi = c] = (αj )δ(yi ,k) pc , where pc = Pr[yi = c] is the prevack lence of class c. The score for a spammer is based on the intuition that the posterior label vector j j (Pr[yi = 1|yi ], . . . , Pr[yi = K|yi ]) for a good annotator will have all the probability mass concentrated on single class. For example for a three class problem (with equal prevalence), a posterior label vector of (1, 0, 0) (certain that the class is one) comes from a good annotator while a (1/3, 1/3, 1/3) (complete uncertainty about the class label) comes from spammer. Based on this they define the following score for each annotator 1 Score = N N K K j j costck Pr[yi = k|yi ]Pr[yi = c|yi ] j i=1 . (10) c=1 k=1 where costck is the misclassification cost when an instance of class c is classified as k. Essentially this is capturing some sort of uncertainty of the posterior label averaged over all the instances. Perfect workers have a score Scorej = 0 while spammers will have high score. An entropic version of this score based on similar ideas has also been recently proposed in [5]. Our proposed spammer score differs from this approach in the following aspects: (1) Implicit in the score defined above (10) j is the assumption that an annotator is a spammer when Pr[yi = c|yi ] = Pr[yi = c], i.e., the estimated posterior labels are simply based on the prevalence and do not depend on the observed labels. By j j Bayes’ rule this is equivalent to Pr[yi |yi = c] = Pr[yi ] which is what we have used to define our spammer score. (2) While both notions of a spammer are equivalent, the approach of [4] first computes the posterior labels based on the observed data, the class prevalence and the annotator j j j j This can be seen as follows: Pr[yi = k | yi = 1] = Pr[(yi ≥ k) AND (yi < k + 1) | yi = 1] = Pr[yi ≥ j j j j k | yi = 1] + Pr[yi < k + 1 | yi = 1] − Pr[(yi ≥ k) OR (yi < k + 1) | yi = 1] = Pr[yi ≥ k | yi = j j j 1] − Pr[yi ≥ k + 1 | yi = 1] = αj − αj . Here we used the fact that Pr[(yi ≥ k) OR (yi < k + 1)] = 1. k k+1 3 5 simulated | 500 instances | 30 annotators simulated | 500 instances | 30 annotators 1 12 0.8 Spammer Score 18 0.6 0.5 22 24 23 25 0.3 29 20 0.2 0.4 0.2 30 16 14 0.1 26 21 27 28 19 0 0 13 0 0.2 0.4 0.6 1−Specificity 0.8 1 500 500 500 500 500 500 500 500 500 500 0.4 0.6 500 500 500 1 0.7 500 500 500 500 500 500 500 500 500 500 500 500 500 500 3 1 500 500 500 2 8 510 7 17 4 9 27 8 30 6 3 28 7 10 2 23 22 26 24 5 1 21 29 25 14 12 17 11 18 20 19 15 16 13 4 0.8 Sensitivity 6 9 0.9 15 11 Annotator (a) Simulation setup (b) Annotator ranking Annotator rank (median) via accuracy simulated | 500 instances | 30 annotators Annotator rank (median) via Ipeirotis et.al.[4] simulated | 500 instances | 30 annotators 27 30 30 28 23 22 26 25 24 21 29 25 20 14 17 18 15 20 16 15 11 13 12 19 1 10 5 10 2 7 6 3 5 8 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (c) Comparison with accuracy 30 18 20 16 15 19 13 12 25 11 14 17 25 20 29 21 51 15 26 22 23 24 10 2 7 10 28 6 3 8 30 5 27 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (d) Comparison with Ipeirotis et. al. [4] Figure 3: (a) The simulation setup consisting of 10 good annotators (annotators 1 to 10), 10 spammers (11 to 20), and 10 malicious annotators (21 to 30). (b) The ranking of annotators obtained using the proposed spammer score. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. (c) and (d) Comparison of the median rank obtained via the spammer score with the rank obtained using (c) accuracy and (d) the method proposed by Ipeirotis et. al. [4]. parameters and then computes the expected cost. Our proposed spammer score does not depend on the prevalence of the class. Our score is also directly defined only in terms of the annotator confusion matrix and does not need the observed labels. (3) For the score defined in (10) while perfect annotators have a score of 0 it is not clear what should be a good baseline for a spammer. The authors suggest to compute the baseline by assuming that a worker assigns as label the class with maximum prevalence. Our proposed score has a natural scale with a perfect annotator having a score of 1 and a spammer having a score of 0. (4) However one advantage of the approach in [4] is that they can directly incorporate varied misclassification costs. 6 Experiments Ranking annotators based on the confidence interval As mentioned earlier the annotator model parameters can be estimated using the iterative EM algorithms [3, 7] and these estimated annotator parameters can then be used to compute the spammer score. The spammer score can then be used to rank the annotators. However one commonly observed phenomenon when working with crowdsourced data is that we have a lot of annotators who label only a very few instances. As a result the annotator parameters cannot be reliably estimated for these annotators. In order to factor this uncertainty in the estimation of the model parameters we compute the spammer score for 100 bootstrap replications. Based on this we compute the 95% confidence intervals (CI) for the spammer score for each annotator. We rank the annotators based on the lower limit of the 95% CI. The CIs are wider 6 Table 1: Datasets N is the number of instances. M is the number of annotators. M ∗ is the mean/median number of annotators per instance. N ∗ is the mean/median number of instances labeled by each annotator. Dataset Type N M M∗ N∗ bluebird binary 108 39 39/39 108/108 temp binary 462 76 10/10 61/16 Brief Description wsd categorical/3 177 34 10/10 52/20 sentiment categorical/3 1660 33 6/6 291/175 30 100 10 38 10/10 10/10 bird identification [12] The annotator had to identify whether there was an Indigo Bunting or Blue Grosbeak in the image. event annotation [10] Given a dialogue and a pair of verbs annotators need to label whether the event described by the first verb occurs before or after the second. 30/30 26/20 30 30 30 30 30 30 3 30 30 1 30 20 20 20 20 20 77 117 20 60 Spammer Score 0.4 10 7 9 8 6 5 0 2 0.2 13 31 10 23 29 1 2 4 6 8 9 14 15 17 22 32 5 18 16 19 11 12 20 21 24 25 26 27 28 30 33 34 7 3 0 0.6 20 20 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 20 20 20 20 17 17 40 20 20 100 Spammer Score 0.4 108 108 108 108 108 108 108 108 108 108 108 108 108 0.8 0.6 0.2 17 8 27 30 25 35 1 12 32 37 38 16 22 9 29 15 20 19 5 39 3 21 23 14 2 10 24 7 33 13 36 31 4 34 28 18 11 6 26 0.2 30 77 77 4 108 108 108 108 0.4 0 wosi | 30 instances | 10 annotators 1 0.8 108 108 0.6 108 108 108 Spammer Score 0.8 wsd | 177 instances | 34 annotators 1 80 177 157 177 157 bluebird | 108 instances | 39 annotators 1 word similarity [10] Numeric judgements of word similarity. affect recognition [10] Each annotator is presented with a short headline and asked to rate it on a scale [-100,100] to denote the overall positive or negative valence. 40 40 20 ordinal/[0 10] ordinal[-100 100] 20 20 20 wosi valence word sense disambiguation [10] The labeler is given a paragraph of text containing the word ”president” and asked to label one of the three appropriate senses. irish economic sentiment analysis [1] Articles from three Irish online news sources were annotated by volunteer users as positive, negative, or irrelevant. 20 20 20 20 20 60 20 20 20 40 40 100 0.4 Annotator Annotator 0 1 26 10 18 28 15 5 36 23 12 8 32 31 38 13 17 27 11 2 35 24 19 9 6 30 33 37 14 29 4 3 20 34 22 25 7 16 21 40 20 0.2 26 2 6 11 5 14 3 20 9 22 31 10 12 18 8 13 30 4 1 29 19 17 27 28 21 15 25 23 7 33 16 24 32 10 132 10 360 10 0 13 18 52 75 33 32 12 74 31 51 41 55 7 14 70 42 58 65 43 1 10 47 61 73 25 37 76 67 24 46 54 48 39 56 15 62 68 44 53 64 40 9 28 6 2 57 3 4 5 8 11 16 17 19 20 21 22 23 26 27 29 30 34 35 36 38 45 49 50 59 60 63 66 69 71 72 442 462 452 10 10 0.6 238 171 75 654 20 0.2 0.2 0 12 77 67 374 249 229 453 346 428 0.4 Spammer Score 43 175 119 541 525 437 0.8 917 104 284 0.4 0.6 1211 1099 10 Spammer Score 0.8 572 30 52 402 60 0.6 30 Spammer Score 0.8 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 60 40 20 15 7 7 11 12 35 29 1 87 10 10 10 10 10 10 10 12 1 30 10 10 10 10 10 10 10 10 10 10 10 10 10 valence | 100 instances | 38 annotators 20 22 10 10 10 10 sentiment | 1660 instances | 33 annotators 10 10 30 20 10 Annotator temp | 462 instances | 76 annotators 1 Annotator 10 50 10 10 40 10 70 350 80 40 100 192 190 40 32 60 70 20 20 40 80 20 50 50 50 30 Annotator Annotator Figure 4: Annotator Rankings The rankings obtained for the datasets in Table 1. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. Note that the CIs are wider when the annotator labels only a few instances. when the annotator labels only a few instances. For a crowdsourced labeling task the annotator has to be good and also label a reasonable number of instances in order to be reliably identified. Simulated data We first illustrate our proposed spammer score on simulated binary data (with equal prevalence for both classes) consisting of 500 instances labeled by 30 annotators of varying sensitivity and specificity (see Figure 3(a) for the simulation setup). Of the 30 annotators we have 10 good annotators (annotators 1 to 10 who lie above the diagonal in Figure 3(a)), 10 spammers (annotators 11 to 20 who lie around the diagonal), and 10 malicious annotators (annotators 21 to 30 who lie below the diagonal). Figure 3(b) plots the ranking of annotators obtained using the proposed spammer score with the annotator model parameters estimated via the EM algorithm [3, 7]. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence interval (CI) obtained via bootstrapping are shown. The annotators are ranked based on the lower limit of the 95% CI. As can be seen all the spammers (annotators 11 to 20) have a low spammer score and appear at the bottom of the list. The malicious annotators have higher score than the spammers since we can correct for their flipping. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we detect that they are malicious. Figure 3(c) compares the (median) rank obtained via the spammer score with the (median) rank obtained using accuracy as the score to rank the annotators. While the good annotators are ranked high by both methods the accuracy score gives a low rank to the malicious annotators. Accuracy does not capture the notion of a spammer. Figure 3(d) compares the ranking with the method proposed by Ipeirotis et. al. [4] which gives almost similar rankings as our proposed score. 7 21 23 10 6 35 4 34 1126 18 147 30 3 31 13 2436 33 25 5 2 20 15 39 19 15 20 28 22 299 12 37 16 38 10 32 1 5 27 25 35 30 8 17 0 0 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 40 bluebird | 108 instances | 39 annotators 40 1 6 34 112618 4 31 1013 7 30 2 28 21 5 20 15 39 19 20 15 22 37 16 299 12 38 10 5 8 17 0 0 27 25 35 30 0.6 0.5 35 32 2 0.4 36 11 13 31 24 10 33 28 21 26 18 0 0 40 34 15 19 39 0.1 (a) 22 37 20 38 29 9 0.2 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 6 4 16 0.3 32 1 7 30 25 1 3 14 3 27 5 0.7 24 33 14 36 23 25 12 8 0.9 17 0.8 35 Sensitivity Annotator rank (median) via accuracy bluebird | 108 instances | 39 annotators Annotator rank (median) via Ipeirotis et.al.[4] bluebird | 108 instances | 39 annotators 40 23 0.2 0.4 0.6 1−Specificity (b) 0.8 1 (c) Figure 5: Comparison of the rank obtained via the spammer score with the rank obtained using (a) accuracy and (b) the method proposed by Ipeirotis et. al. [4] for the bluebird binary dataset. (c) The annotator model parameters as estimated by the EM algorithm [3, 7]. 19 25 12 18 7 3 14 20 32 5 8 1 16 20 9 21 15 34 10 31 29 17 28 22 26 2315 5 2 0 0 4 6 13 10 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 30 25 16 19 7 25 8 9 27 14 3 28 17 18 32 5 10 4 2 10 6 1529 31 23 22 21 15 0 0 33 30 11 1 20 5 sentiment | 1660 instances | 33 annotators 24 35 12 20 24 34 26 13 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 33 7 30 15 17 25 28 2719 2223 20 8 1 4 1812 15 13 10 20 32 30 10 3 29 9 31 16 5 6 2 5 14 11 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score 25 21 Annotator rank (median) via Ipeirotis et.al.[4] 25 24 27 Annotator rank (median) via accuracy Annotator rank (median) via accuracy 30 sentiment | 1660 instances | 33 annotators wsd | 177 instances | 34 annotators 33 30 11 Annotator rank (median) via Ipeirotis et.al.[4] wsd | 177 instances | 34 annotators 35 7 30 15 19 17 27 25 21 25 8 12 4 18 20 24 15 20 33 10 3 13 9 28 1 29 23 10 1632 11 14 5 6 2 5 31 30 22 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score Figure 6: Comparison of the median rank obtained via the spammer score with the rank obtained using accuracy and he method proposed by Ipeirotis et. al. [4] for the two categorial datasets in Table 1. Mechanical Turk data We report results on some publicly available linguistic and image annotation data collected using the Amazon’s Mechanical Turk (AMT) and other sources. Table 1 summarizes the datasets. Figure 4 plots the spammer scores and rankings obtained. The mean and the 95% CI obtained via bootstrapping are also shown. The number at the top of the CI bar shows the number of instances annotated by that annotator. The rankings are based on the lower limit of the 95% CI which factors the number of instances labeled by the annotator into the ranking. An annotator who labels only a few instances will have very wide CI. Some annotators who label only a few instances may have a high mean spammer score but the CI will be wide and hence ranked lower. Ideally we would like to have annotators with a high score and at the same time label a lot of instances so that we can reliablly identify them. The authors [1] for the sentiment dataset shared with us some of the qualitative observations regarding the annotators and they somewhat agree with our rankings. For example the authors made the following comments about Annotator 7 ”Quirky annotator - had a lot of debate about what was the meaning of the annotation question. I’d say he changed his labeling strategy at least once during the process”. Our proposed score gave a low rank to this annotator. Comparison with other approaches Figure 5 and 6 compares the proposed ranking with the rank obtained using accuracy and the method proposed by Ipeirotis et. al. [4] for some binary and categorical datasets in Table 1. Our proposed ranking is somewhat similar to that obtained by Ipeirotis et. al. [4] but accuracy does not quite capture the notion of spammer. For example for the bluebird dataset for annotator 21 (see Figure 5(a)) accuracy ranks it at the bottom of the list while the proposed score puts is in the middle of the list. From the estimated model parameters it can be seen that annotator 21 actually flips the labels (below the diagonal in Figure 5(c)) but is a good annotator. 7 Conclusions We proposed a score to rank annotators for crowdsourced binary, categorical, and ordinal labeling tasks. The obtained rankings and the scores can be used to allocate monetary bonuses to be paid to different annotators and also to eliminate spammers from further labeling tasks. A mechanism to rank annotators should be desirable feature of any crowdsourcing service. The proposed score should also be useful to specify the prior for Bayesian approaches to consolidate annotations. 8 References [1] A. Brew, D. Greene, and P. Cunningham. Using crowdsourcing and active learning to track sentiment in online media. In Proceedings of the 6th Conference on Prestigious Applications of Intelligent Systems (PAIS’10), 2010. [2] B. Carpenter. Multilevel bayesian models of categorical data annotation. Technical Report available at http://lingpipe-blog.com/lingpipe-white-papers/, 2008. [3] A. P. Dawid and A. M. Skene. Maximum likeihood estimation of observer error-rates using the EM algorithm. Applied Statistics, 28(1):20–28, 1979. [4] P. G. Ipeirotis, F. Provost, and J. Wang. Quality management on Amazon Mechanical Turk. In Proceedings of the ACM SIGKDD Workshop on Human Computation (HCOMP’10), pages 64–67, 2010. [5] V. C. Raykar and S. Yu. An entropic score to rank annotators for crowdsourced labelling tasks. In Proceedings of the Third National Conference on Computer Vision, Pattern Recognition, Image Processing and Graphics (NCVPRIPG), 2011. [6] V. C. Raykar, S. Yu, L .H. Zhao, A. Jerebko, C. Florin, G. H. Valadez, L. Bogoni, and L. Moy. Supervised learning from multiple experts: Whom to trust when everyone lies a bit. In Proceedings of the 26th International Conference on Machine Learning (ICML 2009), pages 889– 896, 2009. [7] V. C. Raykar, S. Yu, L. H. Zhao, G. H. Valadez, C. Florin, L. Bogoni, and L. Moy. Learning from crowds. Journal of Machine Learning Research, 11:1297–1322, April 2010. [8] V. S. Sheng, F. Provost, and P. G. Ipeirotis. Get another label? Improving data quality and data mining using multiple, noisy labelers. In Proceedings of the 14th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pages 614–622, 2008. [9] P. Smyth, U. Fayyad, M. Burl, P. Perona, and P. Baldi. Inferring ground truth from subjective labelling of venus images. In Advances in Neural Information Processing Systems 7, pages 1085–1092. 1995. [10] R. Snow, B. O’Connor, D. Jurafsky, and A. Y. Ng. Cheap and Fast—but is it good? Evaluating Non-Expert Annotations for Natural Language Tasks. In Proceedings of the Conference on Empirical Methods in Natural Language Processing (EMNLP ’08), pages 254–263, 2008. [11] A. Sorokin and D. Forsyth. Utility data annotation with Amazon Mechanical Turk. In Proceedings of the First IEEE Workshop on Internet Vision at CVPR 08, pages 1–8, 2008. [12] P. Welinder, S. Branson, S. Belongie, and P. Perona. The multidimensional wisdom of crowds. In Advances in Neural Information Processing Systems 23, pages 2424–2432. 2010. [13] J. Whitehill, P. Ruvolo, T. Wu, J. Bergsma, and J. Movellan. Whose vote should count more: Optimal integration of labels from labelers of unknown expertise. In Advances in Neural Information Processing Systems 22, pages 2035–2043. 2009. [14] Y. Yan, R. Rosales, G. Fung, M. Schmidt, G. Hermosillo, L. Bogoni, L. Moy, and J. Dy. Modeling annotator expertise: Learning when everybody knows a bit of something. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics (AISTATS 2010), pages 932–939, 2010. 9
2 0.063844778 42 nips-2011-Bayesian Bias Mitigation for Crowdsourcing
Author: Fabian L. Wauthier, Michael I. Jordan
Abstract: Biased labelers are a systemic problem in crowdsourcing, and a comprehensive toolbox for handling their responses is still being developed. A typical crowdsourcing application can be divided into three steps: data collection, data curation, and learning. At present these steps are often treated separately. We present Bayesian Bias Mitigation for Crowdsourcing (BBMC), a Bayesian model to unify all three. Most data curation methods account for the effects of labeler bias by modeling all labels as coming from a single latent truth. Our model captures the sources of bias by describing labelers as influenced by shared random effects. This approach can account for more complex bias patterns that arise in ambiguous or hard labeling tasks and allows us to merge data curation and learning into a single computation. Active learning integrates data collection with learning, but is commonly considered infeasible with Gibbs sampling inference. We propose a general approximation strategy for Markov chains to efficiently quantify the effect of a perturbation on the stationary distribution and specialize this approach to active learning. Experiments show BBMC to outperform many common heuristics. 1
3 0.062854417 153 nips-2011-Learning large-margin halfspaces with more malicious noise
Author: Phil Long, Rocco Servedio
Abstract: We describe a simple algorithm that runs in time poly(n, 1/γ, 1/ε) and learns an unknown n-dimensional γ-margin halfspace to accuracy 1 − ε in the presence of malicious noise, when the noise rate is allowed to be as high as Θ(εγ log(1/γ)). Previous efficient algorithms could only learn to accuracy ε in the presence of malicious noise of rate at most Θ(εγ). Our algorithm does not work by optimizing a convex loss function. We show that no algorithm for learning γ-margin halfspaces that minimizes a convex proxy for misclassification error can tolerate malicious noise at a rate greater than Θ(εγ); this may partially explain why previous algorithms could not achieve the higher noise tolerance of our new algorithm. 1
4 0.057461247 137 nips-2011-Iterative Learning for Reliable Crowdsourcing Systems
Author: David R. Karger, Sewoong Oh, Devavrat Shah
Abstract: Crowdsourcing systems, in which tasks are electronically distributed to numerous “information piece-workers”, have emerged as an effective paradigm for humanpowered solving of large scale problems in domains such as image classification, data entry, optical character recognition, recommendation, and proofreading. Because these low-paid workers can be unreliable, nearly all crowdsourcers must devise schemes to increase confidence in their answers, typically by assigning each task multiple times and combining the answers in some way such as majority voting. In this paper, we consider a general model of such crowdsourcing tasks, and pose the problem of minimizing the total price (i.e., number of task assignments) that must be paid to achieve a target overall reliability. We give a new algorithm for deciding which tasks to assign to which workers and for inferring correct answers from the workers’ answers. We show that our algorithm significantly outperforms majority voting and, in fact, is asymptotically optimal through comparison to an oracle that knows the reliability of every worker. 1
5 0.054971375 180 nips-2011-Multiple Instance Filtering
Author: Kamil A. Wnuk, Stefano Soatto
Abstract: We propose a robust filtering approach based on semi-supervised and multiple instance learning (MIL). We assume that the posterior density would be unimodal if not for the effect of outliers that we do not wish to explicitly model. Therefore, we seek for a point estimate at the outset, rather than a generic approximation of the entire posterior. Our approach can be thought of as a combination of standard finite-dimensional filtering (Extended Kalman Filter, or Unscented Filter) with multiple instance learning, whereby the initial condition comes with a putative set of inlier measurements. We show how both the state (regression) and the inlier set (classification) can be estimated iteratively and causally by processing only the current measurement. We illustrate our approach on visual tracking problems whereby the object of interest (target) moves and evolves as a result of occlusions and deformations, and partial knowledge of the target is given in the form of a bounding box (training set). 1
6 0.038551517 66 nips-2011-Crowdclustering
7 0.038305283 40 nips-2011-Automated Refinement of Bayes Networks' Parameters based on Test Ordering Constraints
8 0.037259996 240 nips-2011-Robust Multi-Class Gaussian Process Classification
9 0.036338929 247 nips-2011-Semantic Labeling of 3D Point Clouds for Indoor Scenes
10 0.036270849 159 nips-2011-Learning with the weighted trace-norm under arbitrary sampling distributions
11 0.033898339 22 nips-2011-Active Ranking using Pairwise Comparisons
12 0.032213274 96 nips-2011-Fast and Balanced: Efficient Label Tree Learning for Large Scale Object Recognition
13 0.031953756 6 nips-2011-A Global Structural EM Algorithm for a Model of Cancer Progression
14 0.03110248 303 nips-2011-Video Annotation and Tracking with Active Learning
15 0.02813836 116 nips-2011-Hierarchically Supervised Latent Dirichlet Allocation
16 0.027759276 288 nips-2011-Thinning Measurement Models and Questionnaire Design
17 0.027341092 211 nips-2011-Penalty Decomposition Methods for Rank Minimization
18 0.026364846 193 nips-2011-Object Detection with Grammar Models
19 0.026009766 33 nips-2011-An Exact Algorithm for F-Measure Maximization
20 0.02496111 217 nips-2011-Practical Variational Inference for Neural Networks
topicId topicWeight
[(0, 0.065), (1, 0.014), (2, -0.032), (3, -0.002), (4, 0.013), (5, -0.009), (6, -0.004), (7, -0.043), (8, -0.041), (9, 0.04), (10, 0.015), (11, -0.025), (12, -0.0), (13, -0.017), (14, 0.046), (15, -0.028), (16, -0.03), (17, -0.047), (18, 0.057), (19, -0.04), (20, -0.036), (21, -0.08), (22, 0.026), (23, -0.073), (24, 0.047), (25, 0.012), (26, -0.017), (27, -0.022), (28, -0.0), (29, -0.054), (30, 0.015), (31, -0.027), (32, 0.051), (33, 0.009), (34, -0.007), (35, -0.012), (36, -0.067), (37, 0.04), (38, 0.015), (39, -0.037), (40, -0.01), (41, -0.11), (42, 0.045), (43, 0.029), (44, -0.036), (45, 0.131), (46, 0.067), (47, -0.005), (48, 0.024), (49, 0.15)]
simIndex simValue paperId paperTitle
same-paper 1 0.94925934 232 nips-2011-Ranking annotators for crowdsourced labeling tasks
Author: Vikas C. Raykar, Shipeng Yu
Abstract: With the advent of crowdsourcing services it has become quite cheap and reasonably effective to get a dataset labeled by multiple annotators in a short amount of time. Various methods have been proposed to estimate the consensus labels by correcting for the bias of annotators with different kinds of expertise. Often we have low quality annotators or spammers–annotators who assign labels randomly (e.g., without actually looking at the instance). Spammers can make the cost of acquiring labels very expensive and can potentially degrade the quality of the consensus labels. In this paper we formalize the notion of a spammer and define a score which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a high score close to one. 1 Spammers in crowdsourced labeling tasks Annotating an unlabeled dataset is one of the bottlenecks in using supervised learning to build good predictive models. Getting a dataset labeled by experts can be expensive and time consuming. With the advent of crowdsourcing services (Amazon’s Mechanical Turk being a prime example) it has become quite easy and inexpensive to acquire labels from a large number of annotators in a short amount of time (see [8], [10], and [11] for some computer vision and natural language processing case studies). One drawback of most crowdsourcing services is that we do not have tight control over the quality of the annotators. The annotators can come from a diverse pool including genuine experts, novices, biased annotators, malicious annotators, and spammers. Hence in order to get good quality labels requestors typically get each instance labeled by multiple annotators and these multiple annotations are then consolidated either using a simple majority voting or more sophisticated methods that model and correct for the annotator biases [3, 9, 6, 7, 14] and/or task complexity [2, 13, 12]. In this paper we are interested in ranking annotators based on how spammer like each annotator is. In our context a spammer is a low quality annotator who assigns random labels (maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator). Spammers can significantly increase the cost of acquiring annotations (since they need to be paid) and at the same time decrease the accuracy of the final consensus labels. A mechanism to detect and eliminate spammers is a desirable feature for any crowdsourcing market place. For example one can give monetary bonuses to good annotators and deny payments to spammers. The main contribution of this paper is to formalize the notion of a spammer for binary, categorical, and ordinal labeling tasks. More specifically we define a scalar metric which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a score close to one (see Figure 4). We summarize the multiple parameters corresponding to each annotator into a single score indicative of how spammer like the annotator is. While this spammer score was implicit for binary labels in earlier works [3, 9, 2, 6] the extension to categorical and ordinal labels is novel and is quite different from the accuracy computed from the confusion rate matrix. An attempt to quantify the quality of the workers based on the confusion matrix was recently made by [4] where they transformed the observed labels into posterior soft labels based on the estimated confusion 1 matrix. While we obtain somewhat similar annotator rankings, we differ from this work in that our score is directly defined in terms of the annotator parameters (see § 5 for more details). The rest of the paper is organized as follows. For ease of exposition we start with binary labels (§ 2) and later extend it to categorical (§ 3) and ordinal labels (§ 4). We first specify the annotator model used, formalize the notion of a spammer, and propose an appropriate score in terms of the annotator model parameters. We do not dwell too much on the estimation of the annotator model parameters. These parameters can either be estimated directly using known gold standard 1 or the iterative algorithms that estimate the annotator model parameters without actually knowing the gold standard [3, 9, 2, 6, 7]. In the experimental section (§ 6) we obtain rankings for the annotators using the proposed spammer scores on some publicly available data from different domains. 2 Spammer score for crowdsourced binary labels j Annotator model Let yi ∈ {0, 1} be the label assigned to the ith instance by the j th annotator, and let yi ∈ {0, 1} be the actual (unobserved) binary label. We model the accuracy of the annotator separately on the positive and the negative examples. If the true label is one, the sensitivity (true positive rate) αj for the j th annotator is defined as the probability that the annotator labels it as one. j αj := Pr[yi = 1|yi = 1]. On the other hand, if the true label is zero, the specificity (1−false positive rate) β j is defined as the probability that annotator labels it as zero. j β j := Pr[yi = 0|yi = 0]. Extensions of this basic model have been proposed to include item level difficulty [2, 13] and also to model the annotator performance based on the feature vector [14]. For simplicity we use the basic model proposed in [7] in our formulation. Based on many instances labeled by multiple annotators the maximum likelihood estimator for the annotator parameters (αj , β j ) and also the consensus ground truth (yi ) can be estimated iteratively [3, 7] via the Expectation Maximization (EM) algorithm. The EM algorithm iteratively establishes a particular gold standard (initialized via majority voting), measures the performance of the annotators given that gold standard (M-step), and refines the gold standard based on the performance measures (E-step). Who is a spammer? Intuitively, a spammer assigns labels randomly—maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator. More precisely an annotator is a spammer if the probability j of observed label yi being one given the true label yi is independent of the true label, i.e., j j Pr[yi = 1|yi ] = Pr[yi = 1]. This means that the annotator is assigning labels randomly by flipping a coin with bias without actually looking at the data. Equivalently (1) can be written as j j Pr[yi = 1|yi = 1] = Pr[yi = 1|yi = 0] which implies αj = 1 − β j . (1) j Pr[yi = 1] (2) Hence in the context of the annotator model defined earlier a perfect spammer is an annotator for whom αj + β j − 1 = 0. This corresponds to the diagonal line on the Receiver Operating Characteristic (ROC) plot (see Figure 1(a)) 2 . If αj + β j − 1 < 0 then the annotators lies below the diagonal line and is a malicious annotator who flips the labels. Note that a malicious annotator has discriminatory power if we can detect them and flip their labels. In fact the methods proposed in [3, 7] can automatically flip the labels for the malicious annotators. Hence we define the spammer score for an annotator as S j = (αj + β j − 1)2 (3) An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 1 One of the commonly used strategy to filter out spammers is to inject some items into the annotations with known labels. This is the strategy used by CrowdFlower (http://crowdflower.com/docs/gold). 2 Also note that (αj + β j )/2 is equal to the area shown in the plot and can be considered as a non-parametric approximation to the area under the ROC curve (AUC) based on one observed point. It is also equal to the Balanced Classification Rate (BCR). So a spammer can also be defined as having BCR or AUC equal to 0.5. 2 Equal accuracy contours (prevalence=0.5) 0.9 1 Good Annotators Biased Annotators 0.8 0.9 Sensitivity Sensitivity ( αj ) Spammers 0.5 0.4 j 0.3 j 0.6 0.5 0.4 4 0. 5 0. 3 0. 7 0. 6 0. 4 0. 5 0. 2 0. 3 0. Malicious Annotators 0.2 0.4 0.6 1−Specificity ( βj ) 0.8 0.1 1 0. 0. 2 0. 3 0. 1 0. 0.6 0.5 1 0. 2 0. 2 0. 1 0. 0.4 3 1 0. 0. 2 0. 4 0. 0.2 4 0. 5 0. 0 0 1 3 0. 0.3 0.2 Biased Annotators 4 0.7 6 0. 4 0. 0.8 7 0.3 Area = (α +β )/2 0.2 0 0 0.9 0. 8 0. 8 0. 0.7 6 0. .5 0 5 0. 0.7 [ 1−βj, αj ] 0.6 0.1 6 0. 0.8 0.7 Equal spammer score contours 1 7 0. 8 0. 9 0. Sensitivity 1 (a) Binary annotator model 0.1 1 0. 2 0. 3 0. 0.2 0.4 0.6 1−Specificity 0.8 1 1 0. 0 0 (b) Accuracy 0.2 3 0. 4 0. 0.4 0.6 1−Specificity 5 0. .6 7 0 0. 8 0. 0.8 1 (c) Spammer score Figure 1: (a) For binary labels an annotator is modeled by his/her sensitivity and specificity. A perfect spammer lies on the diagonal line on the ROC plot. (b) Contours of equal accuracy (4) and (c) equal spammer score (3). Accuracy This notion of a spammer is quite different for that of the accuracy of an annotator. An annotator with high accuracy is a good annotator but one with low accuracy is not necessarily a spammer. The accuracy is computed as 1 j Accuracyj = Pr[yi = yi ] = j Pr[yi = 1|yi = k]Pr[yi = k] = αj p + β j (1 − p), (4) k=0 where p := Pr[yi = 1] is the prevalence of the positive class. Note that accuracy depends on prevalence. Our proposed spammer score does not depend on prevalence and essentially quantifies the annotator’s inherent discriminatory power. Figure 1(b) shows the contours of equal accuracy on the ROC plot. Note that annotators below the diagonal line (malicious annotators) have low accuracy. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we can detect them and then correct for the flipping. In fact the EM algorithms [3, 7] can correctly flip the labels for the malicious annotators and hence they should not be treated as spammers. Figure 1(c) also shows the contours of equal score for our proposed score and it can be seen that the malicious annotators have a high score and only annotators along the diagonal have a low score (spammers). Log-odds Another interpretation of a spammer can be seen from the log odds. Using Bayes’ rule the posterior log-odds can be written as log j Pr[yi = 1|yi ] Pr[yi = j 0|yi ] = log j Pr[yi |yi = 1] j Pr[yi |yi = 0] + log p . 1−p Pr[y =1|y j ] p If an annotator is a spammer (i.e., (2) holds) then log Pr[yi =0|yi ] = log 1−p . Essentially the annotator j i i provides no information in updating the posterior log-odds and hence does not contribute to the estimation of the actual true label. 3 Spammer score for categorical labels Annotator model Suppose there are K ≥ 2 categories. We introduce a multinomial parameter αj = (αj , . . . , αj ) for each annotator, where c c1 cK K j αj := Pr[yi = k|yi = c] ck αj = 1. ck and k=1 αj ck The term denotes the probability that annotator j assigns class k to an instance given that the true class is c. When K = 2, αj and αj are sensitivity and specificity, respectively. 11 00 Who is a spammer? As earlier a spammer assigns labels randomly, i.e., j j Pr[yi = k|yi ] = Pr[yi = k], ∀k. 3 j j This is equivalent to Pr[yi = k|yi = c] = Pr[yi = k|yi = c ], ∀c, c , k = 1, . . . , K— which means knowing the true class label being c or c does not change the probability of the annotator’s assigned label. This indicates that the annotator j is a spammer if αj = αj k , ∀c, c , k = 1, . . . , K. ck c (5) Let Aj be the K × K confusion rate matrix with entries [Aj ]ck = αck —a spammer would have 0.50 0.50 0.50 all the rows of Aj equal, for example, Aj = 0.25 0.25 0.25 0.25 0.25 0.25 , for a three class categorical annotation problem. Essentially Aj is a rank one matrix of the form Aj = evj , for some column vector vj ∈ RK that satisfies vj e = 1, where e is column vector of ones. In the binary case we had this natural notion of spammer as an annotator for whom αj + β j − 1 was close to zero. One natural way to summarize (5) would be in terms of the distance (Frobenius norm) of the confusion matrix to the closest rank one approximation, i.e, S j := Aj − eˆj v 2 F, (6) where ˆj solves v ˆj = arg min Aj − evj v vj 2 F s.t. vj e = 1. (7) Solving (7) yields ˆj = (1/K)Aj e, which is the mean of the rows of Aj . Then from (6) we have v Sj = I− 1 ee K 2 Aj = F 1 K (αj − αj k )2 . ck c c < . . . < K. Annotator model It is conceptually easier to think of the true label to be binary, that is, yi ∈ {0, 1}. For example in mammography a lesion is either malignant (1) or benign (0) (which can be confirmed by biopsy) and the BIRADS ordinal scale is a means for the radiologist to quantify the uncertainty based on the digital mammogram. The radiologist assigns a higher value of the label if he/she thinks the true label is closer to one. As earlier we characterize each annotator by the sensitivity and the specificity, but the main difference is that we now define the sensitivity and specificity for j each ordinal label (or threshold) k ∈ {1, . . . , K}. Let αj and βk be the sensitivity and specificity k th respectively of the j annotator corresponding to the threshold k, that is, j j j αj = Pr[yi ≥ k | yi = 1] and βk = Pr[yi < k | yi = 0]. k j j Note that αj = 1, β1 = 0 and αj 1 K+1 = 0, βK+1 = 1 from this definition. Hence each annotator j j is parameterized by a set of 2(K − 1) parameters [αj , β2 , . . . , αj , βK ]. This corresponds to an 2 K empirical ROC curve for the annotator (Figure 2). 4 Who is a spammer? As earlier we define an an1 j k=1 notator j to be a spammer if Pr[yi = k|yi = 1] = 0.9 j k=2 0.8 Pr[yi = k|yi = 0] ∀k = 1, . . . , K. Note that from j 0.7 k=3 [ 1−β , α ] the annotation model we have 3 Pr[yi = k | yi = 0.6 j j j 1] = αk − αk+1 and Pr[yi = k | yi = 0] = 0.5 k=4 j j 0.4 βk+1 − βk . This implies that annotator j is a spam0.3 j j mer if αj − αj k k+1 = βk+1 − βk , ∀k = 1, . . . , K, 0.2 j j 0.1 which leads to αj + βk = αj + β1 = 1, ∀k. This 1 k j 0 0 0.2 0.4 0.6 0.8 1 means that for every k, the point (1 − βk , αj ) lies on k 1−Specificity ( β ) the diagonal line in the ROC plot shown in Figure 2. The area under the empirical ROC curve can be comFigure 2: Ordinal labels: An annotator is modK 1 puted as (see Figure 2) AUCj = 2 k=1 (αj + eled by sensitivity/specificity for each threshold. k+1 j j αj )(βk+1 − βk ), and can be used to define the folk lowing spammer score as (2AUCj − 1)2 to rank the different annotators. 3 Sensitivity ( αj ) 3 j 2 K (αj k+1 k=1 j S = + j αj )(βk+1 k − j βk ) −1 (9) With two levels this expression defaults to the binary case. An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 5 Previous work Recently Ipeirotis et.al. [4] proposed a score for categorical labels based on the expected cost of the posterior label. In this section we briefly describe their approach and compare it with our proposed score. For each instance labeled by the annotator they first compute the posterior (soft) label j j Pr[yi = c|yi ] for c = 1, . . . , K, where yi is the label assigned to the ith instance by the j th annotator and yi is the true unknown label. The posterior label is computed via Bayes’ rule as j j j Pr[yi = c|yi ] ∝ Pr[yi |yi = c]Pr[yi = c] = (αj )δ(yi ,k) pc , where pc = Pr[yi = c] is the prevack lence of class c. The score for a spammer is based on the intuition that the posterior label vector j j (Pr[yi = 1|yi ], . . . , Pr[yi = K|yi ]) for a good annotator will have all the probability mass concentrated on single class. For example for a three class problem (with equal prevalence), a posterior label vector of (1, 0, 0) (certain that the class is one) comes from a good annotator while a (1/3, 1/3, 1/3) (complete uncertainty about the class label) comes from spammer. Based on this they define the following score for each annotator 1 Score = N N K K j j costck Pr[yi = k|yi ]Pr[yi = c|yi ] j i=1 . (10) c=1 k=1 where costck is the misclassification cost when an instance of class c is classified as k. Essentially this is capturing some sort of uncertainty of the posterior label averaged over all the instances. Perfect workers have a score Scorej = 0 while spammers will have high score. An entropic version of this score based on similar ideas has also been recently proposed in [5]. Our proposed spammer score differs from this approach in the following aspects: (1) Implicit in the score defined above (10) j is the assumption that an annotator is a spammer when Pr[yi = c|yi ] = Pr[yi = c], i.e., the estimated posterior labels are simply based on the prevalence and do not depend on the observed labels. By j j Bayes’ rule this is equivalent to Pr[yi |yi = c] = Pr[yi ] which is what we have used to define our spammer score. (2) While both notions of a spammer are equivalent, the approach of [4] first computes the posterior labels based on the observed data, the class prevalence and the annotator j j j j This can be seen as follows: Pr[yi = k | yi = 1] = Pr[(yi ≥ k) AND (yi < k + 1) | yi = 1] = Pr[yi ≥ j j j j k | yi = 1] + Pr[yi < k + 1 | yi = 1] − Pr[(yi ≥ k) OR (yi < k + 1) | yi = 1] = Pr[yi ≥ k | yi = j j j 1] − Pr[yi ≥ k + 1 | yi = 1] = αj − αj . Here we used the fact that Pr[(yi ≥ k) OR (yi < k + 1)] = 1. k k+1 3 5 simulated | 500 instances | 30 annotators simulated | 500 instances | 30 annotators 1 12 0.8 Spammer Score 18 0.6 0.5 22 24 23 25 0.3 29 20 0.2 0.4 0.2 30 16 14 0.1 26 21 27 28 19 0 0 13 0 0.2 0.4 0.6 1−Specificity 0.8 1 500 500 500 500 500 500 500 500 500 500 0.4 0.6 500 500 500 1 0.7 500 500 500 500 500 500 500 500 500 500 500 500 500 500 3 1 500 500 500 2 8 510 7 17 4 9 27 8 30 6 3 28 7 10 2 23 22 26 24 5 1 21 29 25 14 12 17 11 18 20 19 15 16 13 4 0.8 Sensitivity 6 9 0.9 15 11 Annotator (a) Simulation setup (b) Annotator ranking Annotator rank (median) via accuracy simulated | 500 instances | 30 annotators Annotator rank (median) via Ipeirotis et.al.[4] simulated | 500 instances | 30 annotators 27 30 30 28 23 22 26 25 24 21 29 25 20 14 17 18 15 20 16 15 11 13 12 19 1 10 5 10 2 7 6 3 5 8 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (c) Comparison with accuracy 30 18 20 16 15 19 13 12 25 11 14 17 25 20 29 21 51 15 26 22 23 24 10 2 7 10 28 6 3 8 30 5 27 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (d) Comparison with Ipeirotis et. al. [4] Figure 3: (a) The simulation setup consisting of 10 good annotators (annotators 1 to 10), 10 spammers (11 to 20), and 10 malicious annotators (21 to 30). (b) The ranking of annotators obtained using the proposed spammer score. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. (c) and (d) Comparison of the median rank obtained via the spammer score with the rank obtained using (c) accuracy and (d) the method proposed by Ipeirotis et. al. [4]. parameters and then computes the expected cost. Our proposed spammer score does not depend on the prevalence of the class. Our score is also directly defined only in terms of the annotator confusion matrix and does not need the observed labels. (3) For the score defined in (10) while perfect annotators have a score of 0 it is not clear what should be a good baseline for a spammer. The authors suggest to compute the baseline by assuming that a worker assigns as label the class with maximum prevalence. Our proposed score has a natural scale with a perfect annotator having a score of 1 and a spammer having a score of 0. (4) However one advantage of the approach in [4] is that they can directly incorporate varied misclassification costs. 6 Experiments Ranking annotators based on the confidence interval As mentioned earlier the annotator model parameters can be estimated using the iterative EM algorithms [3, 7] and these estimated annotator parameters can then be used to compute the spammer score. The spammer score can then be used to rank the annotators. However one commonly observed phenomenon when working with crowdsourced data is that we have a lot of annotators who label only a very few instances. As a result the annotator parameters cannot be reliably estimated for these annotators. In order to factor this uncertainty in the estimation of the model parameters we compute the spammer score for 100 bootstrap replications. Based on this we compute the 95% confidence intervals (CI) for the spammer score for each annotator. We rank the annotators based on the lower limit of the 95% CI. The CIs are wider 6 Table 1: Datasets N is the number of instances. M is the number of annotators. M ∗ is the mean/median number of annotators per instance. N ∗ is the mean/median number of instances labeled by each annotator. Dataset Type N M M∗ N∗ bluebird binary 108 39 39/39 108/108 temp binary 462 76 10/10 61/16 Brief Description wsd categorical/3 177 34 10/10 52/20 sentiment categorical/3 1660 33 6/6 291/175 30 100 10 38 10/10 10/10 bird identification [12] The annotator had to identify whether there was an Indigo Bunting or Blue Grosbeak in the image. event annotation [10] Given a dialogue and a pair of verbs annotators need to label whether the event described by the first verb occurs before or after the second. 30/30 26/20 30 30 30 30 30 30 3 30 30 1 30 20 20 20 20 20 77 117 20 60 Spammer Score 0.4 10 7 9 8 6 5 0 2 0.2 13 31 10 23 29 1 2 4 6 8 9 14 15 17 22 32 5 18 16 19 11 12 20 21 24 25 26 27 28 30 33 34 7 3 0 0.6 20 20 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 20 20 20 20 17 17 40 20 20 100 Spammer Score 0.4 108 108 108 108 108 108 108 108 108 108 108 108 108 0.8 0.6 0.2 17 8 27 30 25 35 1 12 32 37 38 16 22 9 29 15 20 19 5 39 3 21 23 14 2 10 24 7 33 13 36 31 4 34 28 18 11 6 26 0.2 30 77 77 4 108 108 108 108 0.4 0 wosi | 30 instances | 10 annotators 1 0.8 108 108 0.6 108 108 108 Spammer Score 0.8 wsd | 177 instances | 34 annotators 1 80 177 157 177 157 bluebird | 108 instances | 39 annotators 1 word similarity [10] Numeric judgements of word similarity. affect recognition [10] Each annotator is presented with a short headline and asked to rate it on a scale [-100,100] to denote the overall positive or negative valence. 40 40 20 ordinal/[0 10] ordinal[-100 100] 20 20 20 wosi valence word sense disambiguation [10] The labeler is given a paragraph of text containing the word ”president” and asked to label one of the three appropriate senses. irish economic sentiment analysis [1] Articles from three Irish online news sources were annotated by volunteer users as positive, negative, or irrelevant. 20 20 20 20 20 60 20 20 20 40 40 100 0.4 Annotator Annotator 0 1 26 10 18 28 15 5 36 23 12 8 32 31 38 13 17 27 11 2 35 24 19 9 6 30 33 37 14 29 4 3 20 34 22 25 7 16 21 40 20 0.2 26 2 6 11 5 14 3 20 9 22 31 10 12 18 8 13 30 4 1 29 19 17 27 28 21 15 25 23 7 33 16 24 32 10 132 10 360 10 0 13 18 52 75 33 32 12 74 31 51 41 55 7 14 70 42 58 65 43 1 10 47 61 73 25 37 76 67 24 46 54 48 39 56 15 62 68 44 53 64 40 9 28 6 2 57 3 4 5 8 11 16 17 19 20 21 22 23 26 27 29 30 34 35 36 38 45 49 50 59 60 63 66 69 71 72 442 462 452 10 10 0.6 238 171 75 654 20 0.2 0.2 0 12 77 67 374 249 229 453 346 428 0.4 Spammer Score 43 175 119 541 525 437 0.8 917 104 284 0.4 0.6 1211 1099 10 Spammer Score 0.8 572 30 52 402 60 0.6 30 Spammer Score 0.8 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 60 40 20 15 7 7 11 12 35 29 1 87 10 10 10 10 10 10 10 12 1 30 10 10 10 10 10 10 10 10 10 10 10 10 10 valence | 100 instances | 38 annotators 20 22 10 10 10 10 sentiment | 1660 instances | 33 annotators 10 10 30 20 10 Annotator temp | 462 instances | 76 annotators 1 Annotator 10 50 10 10 40 10 70 350 80 40 100 192 190 40 32 60 70 20 20 40 80 20 50 50 50 30 Annotator Annotator Figure 4: Annotator Rankings The rankings obtained for the datasets in Table 1. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. Note that the CIs are wider when the annotator labels only a few instances. when the annotator labels only a few instances. For a crowdsourced labeling task the annotator has to be good and also label a reasonable number of instances in order to be reliably identified. Simulated data We first illustrate our proposed spammer score on simulated binary data (with equal prevalence for both classes) consisting of 500 instances labeled by 30 annotators of varying sensitivity and specificity (see Figure 3(a) for the simulation setup). Of the 30 annotators we have 10 good annotators (annotators 1 to 10 who lie above the diagonal in Figure 3(a)), 10 spammers (annotators 11 to 20 who lie around the diagonal), and 10 malicious annotators (annotators 21 to 30 who lie below the diagonal). Figure 3(b) plots the ranking of annotators obtained using the proposed spammer score with the annotator model parameters estimated via the EM algorithm [3, 7]. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence interval (CI) obtained via bootstrapping are shown. The annotators are ranked based on the lower limit of the 95% CI. As can be seen all the spammers (annotators 11 to 20) have a low spammer score and appear at the bottom of the list. The malicious annotators have higher score than the spammers since we can correct for their flipping. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we detect that they are malicious. Figure 3(c) compares the (median) rank obtained via the spammer score with the (median) rank obtained using accuracy as the score to rank the annotators. While the good annotators are ranked high by both methods the accuracy score gives a low rank to the malicious annotators. Accuracy does not capture the notion of a spammer. Figure 3(d) compares the ranking with the method proposed by Ipeirotis et. al. [4] which gives almost similar rankings as our proposed score. 7 21 23 10 6 35 4 34 1126 18 147 30 3 31 13 2436 33 25 5 2 20 15 39 19 15 20 28 22 299 12 37 16 38 10 32 1 5 27 25 35 30 8 17 0 0 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 40 bluebird | 108 instances | 39 annotators 40 1 6 34 112618 4 31 1013 7 30 2 28 21 5 20 15 39 19 20 15 22 37 16 299 12 38 10 5 8 17 0 0 27 25 35 30 0.6 0.5 35 32 2 0.4 36 11 13 31 24 10 33 28 21 26 18 0 0 40 34 15 19 39 0.1 (a) 22 37 20 38 29 9 0.2 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 6 4 16 0.3 32 1 7 30 25 1 3 14 3 27 5 0.7 24 33 14 36 23 25 12 8 0.9 17 0.8 35 Sensitivity Annotator rank (median) via accuracy bluebird | 108 instances | 39 annotators Annotator rank (median) via Ipeirotis et.al.[4] bluebird | 108 instances | 39 annotators 40 23 0.2 0.4 0.6 1−Specificity (b) 0.8 1 (c) Figure 5: Comparison of the rank obtained via the spammer score with the rank obtained using (a) accuracy and (b) the method proposed by Ipeirotis et. al. [4] for the bluebird binary dataset. (c) The annotator model parameters as estimated by the EM algorithm [3, 7]. 19 25 12 18 7 3 14 20 32 5 8 1 16 20 9 21 15 34 10 31 29 17 28 22 26 2315 5 2 0 0 4 6 13 10 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 30 25 16 19 7 25 8 9 27 14 3 28 17 18 32 5 10 4 2 10 6 1529 31 23 22 21 15 0 0 33 30 11 1 20 5 sentiment | 1660 instances | 33 annotators 24 35 12 20 24 34 26 13 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 33 7 30 15 17 25 28 2719 2223 20 8 1 4 1812 15 13 10 20 32 30 10 3 29 9 31 16 5 6 2 5 14 11 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score 25 21 Annotator rank (median) via Ipeirotis et.al.[4] 25 24 27 Annotator rank (median) via accuracy Annotator rank (median) via accuracy 30 sentiment | 1660 instances | 33 annotators wsd | 177 instances | 34 annotators 33 30 11 Annotator rank (median) via Ipeirotis et.al.[4] wsd | 177 instances | 34 annotators 35 7 30 15 19 17 27 25 21 25 8 12 4 18 20 24 15 20 33 10 3 13 9 28 1 29 23 10 1632 11 14 5 6 2 5 31 30 22 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score Figure 6: Comparison of the median rank obtained via the spammer score with the rank obtained using accuracy and he method proposed by Ipeirotis et. al. [4] for the two categorial datasets in Table 1. Mechanical Turk data We report results on some publicly available linguistic and image annotation data collected using the Amazon’s Mechanical Turk (AMT) and other sources. Table 1 summarizes the datasets. Figure 4 plots the spammer scores and rankings obtained. The mean and the 95% CI obtained via bootstrapping are also shown. The number at the top of the CI bar shows the number of instances annotated by that annotator. The rankings are based on the lower limit of the 95% CI which factors the number of instances labeled by the annotator into the ranking. An annotator who labels only a few instances will have very wide CI. Some annotators who label only a few instances may have a high mean spammer score but the CI will be wide and hence ranked lower. Ideally we would like to have annotators with a high score and at the same time label a lot of instances so that we can reliablly identify them. The authors [1] for the sentiment dataset shared with us some of the qualitative observations regarding the annotators and they somewhat agree with our rankings. For example the authors made the following comments about Annotator 7 ”Quirky annotator - had a lot of debate about what was the meaning of the annotation question. I’d say he changed his labeling strategy at least once during the process”. Our proposed score gave a low rank to this annotator. Comparison with other approaches Figure 5 and 6 compares the proposed ranking with the rank obtained using accuracy and the method proposed by Ipeirotis et. al. [4] for some binary and categorical datasets in Table 1. Our proposed ranking is somewhat similar to that obtained by Ipeirotis et. al. [4] but accuracy does not quite capture the notion of spammer. For example for the bluebird dataset for annotator 21 (see Figure 5(a)) accuracy ranks it at the bottom of the list while the proposed score puts is in the middle of the list. From the estimated model parameters it can be seen that annotator 21 actually flips the labels (below the diagonal in Figure 5(c)) but is a good annotator. 7 Conclusions We proposed a score to rank annotators for crowdsourced binary, categorical, and ordinal labeling tasks. The obtained rankings and the scores can be used to allocate monetary bonuses to be paid to different annotators and also to eliminate spammers from further labeling tasks. A mechanism to rank annotators should be desirable feature of any crowdsourcing service. The proposed score should also be useful to specify the prior for Bayesian approaches to consolidate annotations. 8 References [1] A. Brew, D. Greene, and P. Cunningham. Using crowdsourcing and active learning to track sentiment in online media. In Proceedings of the 6th Conference on Prestigious Applications of Intelligent Systems (PAIS’10), 2010. [2] B. Carpenter. Multilevel bayesian models of categorical data annotation. Technical Report available at http://lingpipe-blog.com/lingpipe-white-papers/, 2008. [3] A. P. Dawid and A. M. Skene. Maximum likeihood estimation of observer error-rates using the EM algorithm. Applied Statistics, 28(1):20–28, 1979. [4] P. G. Ipeirotis, F. Provost, and J. Wang. Quality management on Amazon Mechanical Turk. In Proceedings of the ACM SIGKDD Workshop on Human Computation (HCOMP’10), pages 64–67, 2010. [5] V. C. Raykar and S. Yu. An entropic score to rank annotators for crowdsourced labelling tasks. In Proceedings of the Third National Conference on Computer Vision, Pattern Recognition, Image Processing and Graphics (NCVPRIPG), 2011. [6] V. C. Raykar, S. Yu, L .H. Zhao, A. Jerebko, C. Florin, G. H. Valadez, L. Bogoni, and L. Moy. Supervised learning from multiple experts: Whom to trust when everyone lies a bit. In Proceedings of the 26th International Conference on Machine Learning (ICML 2009), pages 889– 896, 2009. [7] V. C. Raykar, S. Yu, L. H. Zhao, G. H. Valadez, C. Florin, L. Bogoni, and L. Moy. Learning from crowds. Journal of Machine Learning Research, 11:1297–1322, April 2010. [8] V. S. Sheng, F. Provost, and P. G. Ipeirotis. Get another label? Improving data quality and data mining using multiple, noisy labelers. In Proceedings of the 14th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pages 614–622, 2008. [9] P. Smyth, U. Fayyad, M. Burl, P. Perona, and P. Baldi. Inferring ground truth from subjective labelling of venus images. In Advances in Neural Information Processing Systems 7, pages 1085–1092. 1995. [10] R. Snow, B. O’Connor, D. Jurafsky, and A. Y. Ng. Cheap and Fast—but is it good? Evaluating Non-Expert Annotations for Natural Language Tasks. In Proceedings of the Conference on Empirical Methods in Natural Language Processing (EMNLP ’08), pages 254–263, 2008. [11] A. Sorokin and D. Forsyth. Utility data annotation with Amazon Mechanical Turk. In Proceedings of the First IEEE Workshop on Internet Vision at CVPR 08, pages 1–8, 2008. [12] P. Welinder, S. Branson, S. Belongie, and P. Perona. The multidimensional wisdom of crowds. In Advances in Neural Information Processing Systems 23, pages 2424–2432. 2010. [13] J. Whitehill, P. Ruvolo, T. Wu, J. Bergsma, and J. Movellan. Whose vote should count more: Optimal integration of labels from labelers of unknown expertise. In Advances in Neural Information Processing Systems 22, pages 2035–2043. 2009. [14] Y. Yan, R. Rosales, G. Fung, M. Schmidt, G. Hermosillo, L. Bogoni, L. Moy, and J. Dy. Modeling annotator expertise: Learning when everybody knows a bit of something. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics (AISTATS 2010), pages 932–939, 2010. 9
2 0.49937704 40 nips-2011-Automated Refinement of Bayes Networks' Parameters based on Test Ordering Constraints
Author: Omar Z. Khan, Pascal Poupart, John-mark M. Agosta
Abstract: In this paper, we derive a method to refine a Bayes network diagnostic model by exploiting constraints implied by expert decisions on test ordering. At each step, the expert executes an evidence gathering test, which suggests the test’s relative diagnostic value. We demonstrate that consistency with an expert’s test selection leads to non-convex constraints on the model parameters. We incorporate these constraints by augmenting the network with nodes that represent the constraint likelihoods. Gibbs sampling, stochastic hill climbing and greedy search algorithms are proposed to find a MAP estimate that takes into account test ordering constraints and any data available. We demonstrate our approach on diagnostic sessions from a manufacturing scenario. 1 INTRODUCTION The problem of learning-by-example has the promise to create strong models from a restricted number of cases; certainly humans show the ability to generalize from limited experience. Machine Learning has seen numerous approaches to learning task performance by imitation, going back to some of the approaches to inductive learning from examples [14]. Of particular interest are problemsolving tasks that use a model to infer the source, or cause of a problem from a sequence of investigatory steps or tests. The specific example we adopt is a diagnostic task such as appears in medicine, electro-mechanical fault isolation, customer support and network diagnostics, among others. We define a diagnostic sequence as consisting of the assignment of values to a subset of tests. The diagnostic process embodies the choice of the best next test to execute at each step in the sequence, by measuring the diagnostic value among the set of available tests at each step, that is, the ability of a test to distinguish among the possible causes. One possible implementation with which to carry out this process, the one we apply, is a Bayes network [9]. As with all model-based approaches, provisioning an adequate model can be daunting, resulting in a “knowledge elicitation bottleneck.” A recent approach for easing the bottleneck grew out of the realization that the best time to gain an expert’s insight into the model structure is during the diagnostic process. Recent work in “QueryBased Diagnostics” [1] demonstrated a way to improve model quality by merging model use and model building into a single process. More precisely the expert can take steps to modify the network structure to add or remove nodes or links, interspersed within the diagnostic sequence. In this paper we show how to extend this variety of learning-by-example to include also refinement of model parameters based on the expert’s choice of test, from which we determine constraints. The nature of these constraints, as shown herein, is derived from the value of the tests to distinguish causes, a value referred to informally as value of information [10]. It is the effect of these novel constraints on network parameter learning that is elucidated in this paper. ∗ J. M. Agosta is no longer affiliated with Intel Corporation 1 Conventional statistical learning approaches are not suited to this problem, since the number of cases available from diagnostic sessions is small, and the data from any case is sparse. (Only a fraction of the tests are taken.) But more relevant is that one diagnostic sequence from an expert user represents the true behavior expected of the model, rather than a noisy realization of a case generated by the true model. We adopt a Bayesian approach, which offers a principled way to incorporate knowledge (constraints and data, when available) and also consider weakening the constraints, by applying a likelihood to them, so that possibly conflicting constraints can be incorporated consistently. Sec. 2 reviews related work and Sec. 3 provides some background on diagnostic networks and model consistency. Then, Sec. 4 describes an augmented Bayesian network that incorporates constraints implied by an expert’s choice of tests. Some sampling techniques are proposed to find the Maximum a posterior setting of the parameters given the constraints (and any data available). The approach is evaluated in Sec. 5 on synthetic data and a real world manufacturing diagnostic scenario. Finally, Sec. 6 discusses some future work. 2 RELATED WORK Parameter learning for Bayesian networks can be viewed as searching in a high-dimensional space. Adopting constraints on the parameters based on some domain knowledge is a way of pruning this search space and learning the parameters more efficiently, both in terms of data needed and time required. Qualitative probabilistic networks [17] allow qualitative constraints on the parameter space to be specified by experts. For instance, the influence of one variable on another, or the combined influence of multiple variables on another variable [5] leads to linear inequalities on the parameters. Wittig and Jameson [18] explain how to transform the likelihood of violating qualitative constraints into a penalty term to adjust maximum likelihood, which allows gradient ascent and Expectation Maximization (EM) to take into account linear qualitative constraints. Other examples of qualitative constraints include some parameters being larger than others, bounded in a range, within ϵ of each other, etc. Various proposals have been made that exploit such constraints. Altendorf et al. [2] provide an approximate technique based on constrained convex optimization for parameter learning. Niculescu et al. [15] also provide a technique based on constrained optimization with closed form solutions for different classes of constraints. Feelders [6] provides an alternate method based on isotonic regression while Liao and Ji [12] combine gradient descent with EM. de Campos and Ji [4] also use constrained convex optimization, however, they use Dirichlet priors on the parameters to incorporate any additional knowledge. Mao and Lebanon [13] also use Dirichlet priors, but they use probabilistic constraints to allow inaccuracies in the specification of the constraints. A major difference between our technique and previous work is on the type of constraints. Our constraints do not need to be explicitly specified by an expert. Instead, we passively observe the expert and learn from what choices are made and not made [16]. Furthermore, as we shall show later, our constraints are non-convex, preventing the direct application of existing techniques that assume linear or convex functions. We use Beta priors on the parameters, which can easily be extended to Dirichlet priors like previous work. We incorporate constraints in an augmented Bayesian network, similar to Liang et al. [11], though their constraints are on model predictions as opposed to ours which are on the parameters of the network. Finally, we also use the notion of probabilistic constraints to handle potential mistakes made by experts. 3 3.1 BACKGROUND DIAGNOSTIC BAYES NETWORKS We consider the class of bipartite Bayes networks that are widely used as diagnostic models, though our approach can be used for networks with any structure. The network forms a sparse, directed, causal graph, where arcs go from causes to observable node variables. We use upper case to denote random variables; C for causes, and T for observables (tests). Lower case letters denote values in the domain of a variable, e.g. c ∈ dom(C) = {c, c}, and bold letters denote sets of variables. A ¯ set of marginally independent binary-valued node variables C with distributions Pr(C) represent unobserved causes, and condition the remaining conditionally independent binary-valued test vari2 able nodes T. Each cause conditions one or more tests; likewise each test is conditioned by one or more causes, resulting in a graph with one or more possibly multiply-connected components. The test variable distributions Pr(T |C) incorporate the further modeling assumption of Independence of Causal Influence, the most familiar example being the Noisy-Or model [8]. To keep the exposition simple, we assume that all variables are binary and that conditional distributions are parametrized by the Noisy-Or; however, the algorithms described in the rest of the paper generalize to any discrete non-binary variable models. Conventionally, unobserved tests are ranked in a diagnostic Bayes network by their Value Of Information (VOI) conditioned on tests already observed. To be precise, VOI is the expected gain in utility if the test were to be observed. The complete computation requires a model equivalent to a partially observable Markov decision process. Instead, VOI is commonly approximated by a greedy computation of the Mutual Information between a test and the set of causes [3]. In this case, it is easy to show that Mutual Information is in turn well approximated to second order by the Gini impurity [7] as shown in Equation 1. ] [∑ ∑ GI(C|T ) = Pr(T = t) Pr(C = c|T = t)(1 − Pr(C = c|T = t)) (1) t c We will use the Gini measure as a surrogate for VOI, as a way to rank the best next test in the diagnostic sequence. 3.2 MODEL CONSISTENCY A model that is consistent with an expert would generate Gini impurity rankings consistent with the expert’s diagnostic sequence. We interpret the expert’s test choices as implying constraints on Gini impurity rankings between tests. To that effect, [1] defines the notion of Cause Consistency and Test Consistency, which indicate whether the cause and test orderings induced by the posterior distribution over causes and the VOI of each test agree with an expert’s observed choice. Assuming that the expert greedily chooses the most informative test T ∗ (i.e., test that yields the lowest Gini impurity) at each step, then the model is consistent with the expert’s choices when the following constraints are satisfied: GI(C|T ∗ ) ≤ GI(C|Ti ) ∀i (2) We demonstrate next how to exploit these constraints to refine the Bayes network. 4 MODEL REFINEMENT Consider a simple diagnosis example with two possible causes C1 and C2 and two tests T1 and T2 as shown in Figure 1. To keep the exposition simple, suppose that the priors for each cause are known (generally separate data is available to estimate these), but the conditional distribution of each test is unknown. Using the Noisy-OR parameterizations for the conditional distributions, the number of parameters are linear in the number of parents instead of exponential. ∏ i i Pr(Ti = true|C) = 1 − (1 − θ0 ) (1 − θj ) (3) j|Cj =true i Here, θ0 = Pr(Ti = true|Cj = f alse ∀j) is the leak probability that Ti will be true when none of i the causes are true and θj = Pr(Ti = true|Cj = true, Ck = f alse ∀k ̸= j) is the link reliability, which indicates the independent contribution of cause Cj to the probability that test Ti will be true. In the rest of this section, we describe how to learn the θ parameters while respecting the constraints implied by test consistency. 4.1 TEST CONSISTENCY CONSTRAINTS Suppose that an expert chooses test T1 instead of test T2 during the diagnostic process. This ordering by the expert implies that the current model (parametrized by the θ’s) must be consistent with the constraint GI(C|T2 ) − GI(C|T1 ) ≥ 0. Using the definition of Gini impurity in Eq. 1, we can rewrite 3 Figure 1: Network with 2 causes and 2 tests Figure 2: Augmented network with parameters and constraints Figure 3: Augmented network extended to handle inaccurate feedback the constraint for the network shown in Fig. 1 as follows: ∑ t1 ( ∑ (Pr(t1 |c1 , c2 ) Pr(c1 ) Pr(c2 ))2 Pr(t1 ) − Pr(t1 ) c ,c 1 2 ) ( ) ∑ ∑ (Pr(t2 |c1 , c2 ) Pr(c1 ) Pr(c2 ))2 − Pr(t2 ) − ≥0 Pr(t2 ) t c ,c 2 1 2 (4) Furthermore, using the Noisy-Or encoding from Eq. 3, we can rewrite the constraint as a polynomial in the θ’s. This polynomial is non-linear, and in general, not concave. The feasible space may consist of disconnected regions. Fig. 4 shows the surface corresponding to the polynomial for the 2 1 i i case where θ0 = 0 and θ1 = 0.5 for each test i, which leaves θ2 and θ2 as the only free variables. The parameters’ feasible space, satisfying the constraint consists of the two disconnected regions where the surface is positive. 4.2 AUGMENTED BAYES NETWORK Our objective is to learn the θ parameters of diagnostic Bayes networks given test constraints of the form described in Eq. 4. To deal with non-convex constraints and disconnected feasible regions, we pursue a Bayesian approach whereby we explicitly model the parameters and constraints as random variables in an augmented Bayes network (see Fig. 2). This allows us to frame the problem of learning the parameters as an inference problem in a hybrid Bayes network of discrete (T, C, V ) and continuous (Θ) variables. As we will see shortly, this augmented Bayes network provides a unifying framework to simultaneously learn from constraints and data, to deal with possibly inconsistent constraints, and to express preferences over the degree of satisfaction of the constraints. We encode the constraint derived from the expert feedback as a binary random variable V in the Bayes network. If V is true the constraint is satisfied, otherwise it is violated. Thus, if V is true then Θ lies in the positive region of Fig. 4, and if V is f alse then Θ lies in the negative region. We model the CPT for V as Pr(V |Θ) = max(0, π), where π = GI(C|T1 ) − GI(C|T2 ). Note that the value of GI(C|T ) lies in the interval [0,1], so the probability π will always be normalized. The intuition behind this definition of the CPT for V is that a constraint is more likely to be satisfied if the parameters lie in the interior of the constraint region. We place a Beta prior over each Θ parameter. Since the test variables are conditioned on the Θ parameters that are now part of the network, their conditional distributions become known. For instance, the conditional distribution for Ti (given in Eq. 3) is fully defined given the noisy-or parami eters θj . Hence the problem of learning the parameters becomes an inference problem to compute posteriors over the parameters given that the constraint is satisfied (and any data). In practice, it is more convenient to obtain a single value for the parameters instead of a posterior distribution since it is easier to make diagnostic predictions based on one Bayes network. We estimate the parameters by computing a maximum a posteriori (MAP) hypothesis given that the constraint is satisfied (and any data): Θ∗ = arg maxΘ Pr(Θ|V = true). 4 Algorithm 1 Pseudo Code for Gibbs Sampling, Stochastic Hill Climbing and Greedy Search 1 Fix observed variables, let V = true and randomly sample feasible starting state S 2 for i = 1 to #samples 3 for j = 1 to #hiddenV ariables 4 acceptSample = f alse; k = 0 5 repeat 6 Sample s′ from conditional of j th hidden variable Sj 7 S′ = S; Sj = s′ 8 if Sj is cause or test, then acceptSample = true 9 elseif S′ obeys constraints V∗ 10 if algo == Gibbs 11 Sample u from uniform distribution, U(0,1) p(S′ 12 if u < M q(S)′ ) where p and q are the true and proposal distributions and M > 1 13 acceptSample = true 14 elseif algo = = StochasticHillClimbing 15 if likelihood(S′ ) > likelihood(S), then acceptSample = true 16 elseif algo = = Greedy, then acceptSample = true 17 elseif algo = = Greedy 18 k = k+1 19 if k = = maxIterations, then s′ = Sj ; acceptSample = true 20 until acceptSample = = true 21 Sj = s′ 4.3 MAP ESTIMATION Previous approaches for parameter learning with domain knowledge include modified versions of EM or some other optimization techniques that account for linear/convex constraints on the parameters. Since our constraints are non-convex, we propose a new approach based on Gibbs sampling to approximate the posterior distribution, from which we compute the MAP estimate. Although the technique converges to the MAP in the limit, it may require excessive time. Hence, we modify Gibbs sampling to obtain more efficient stochastic hill climbing and greedy search algorithms with anytime properties. The pseudo code for our Gibbs sampler is provided in Algorithm 1. The two key steps are sampling the conditional distributions of each variable (line 6) and rejection sampling to ensure that the constraints are satisfied (lines 9 and 12). We sample each variable given the rest according to the following distributions: ti ∼ Pr(Ti |c, θi ) ∀i cj ∼ Pr(Cj |c − cj , t, θ) ∝ ∏ Pr(Cj ) j ∏ (5) Pr(ti |c, θi ) ∀j i i θj ∼ Pr(Θi |Θ − Θi , t, c, v) ∝ Pr(v|t, Θ) j j ∏ Pr(ti |cj , θi ) ∀i, j (6) (7) i The tests and causes are easily sampled from the multinomials as described in the equations above. However, sampling the θ’s is more difficult due to the factor Pr(v|Θ, t) = max(0, π), which is a truncated mixture of Betas. So, instead of sampling θ from its true conditional, we sample it from a proposal distribution that replaces max(0, π) by an un-truncated mixture of Betas equal to π + a where a is a constant that ensures that π + a is always positive. This is equivalent to ignoring the constraints. Then we ensure that the constraints are satisfied by rejecting the samples that violate the constraints. Once Gibbs sampling has been performed, we obtain a sample that approximates the posterior distribution over the parameters given the constraints (and any data). We return a single setting of the parameters by selecting the sampled instance with the highest posterior probability (i.e., MAP estimate). Since we will only return the MAP estimate, it is possible to speed up the search by modifying Gibbs sampling. In particular, we obtain a stochastic hill climbing algorithm by accepting a new sample only if its posterior probability improves upon that of the previous sample 5 Posterior Probability 0.1 0.08 Difference in Gini Impurity 0.1 0.05 0 −0.05 0.06 0.04 0.02 −0.1 1 0 1 1 0.8 0.5 0.6 0.8 0.4 Link Reliability of Test 2 and Cause 1 0 0.6 0.2 0 0.4 Link Reliability of Test 2 and Cause 2 Figure 4: Difference in Gini impurity for the network in 1 2 Fig. 1 when θ2 and θ2 are the only parameters allowed to vary. 0.2 Link Reliability of Test 2 and Cause 1 0 0 0.2 0.4 0.6 0.8 1 Link Reliability of Test 2 and Cause 1 Figure 5: Posterior over parameters computed through calculation after discretization. Figure 6: Posterior over parameters calculated through Sampling. (line 15). Thus, each iteration of the stochastic hill climber requires more time, but always improves the solution. As the number of constraints grows and the feasibility region shrinks, the Gibbs sampler and stochastic hill climber will reject most samples. We can mitigate this by using a Greedy sampler that caps the number of rejected samples, after which it abandons the sampling for the current variable to move on to the next variable (line 19). Even though the feasibility region is small overall, it may still be large in some dimensions, so it makes sense to try sampling another variable (that may have a larger range of feasible values) when it is taking too long to find a new feasible value for the current variable. 4.4 MODEL REFINEMENT WITH INCONSISTENT CONSTRAINTS So far, we have assumed that the expert’s actions generate a feasible region as a consequence of consistent constraints. We handle inconsistencies by further extending our augmented diagnostic Bayes network. We treat the observed constraint variable, V , as a probabilistic indicator of the true constraint V ∗ as shown in Figure 3. We can easily extend our techniques for computing the MAP to cater for this new constraint node by sampling an extra variable. 5 EVALUATION AND EXPERIMENTS 5.1 EVALUATION CRITERIA Formally, for M ∗ , the true model that we aim to learn, the diagnostic process determines the choice of best next test as the one with the smallest Gini impurity. If the correct choice for the next test is known (such as demonstrated by an expert), we can use this information to include a constraint on the model. We denote by V+ the set of observed constraints and by V∗ the set of all possible constraints that hold for M ∗ . Having only observed V+ , our technique will consider any M + ∈ M+ as a possible true model, where M+ is the set of all models that obey V + . We denote by M∗ the set of all models that are diagnostically equivalent to M ∗ (i.e., obey V ∗ and would recommend the MAP same steps as M ∗ ) and by MV+ the particular model obtained by MAP estimation based on the MAP constraints V+ . Similarly, when a dataset D is available, we denote by MD the model obtained MAP by MAP estimation based on D and by MDV+ , the model based on D and V+ . Ideally we would like to find the true underlying model M ∗ , hence we will report the KL divergence between the models found and M ∗ . However, other diagnostically equivalent M ∗ may recommend the same tests as M ∗ and thus have similar constraints, so we also report test consistency with M ∗ (i.e., # of recommended tests that are the same). 5.2 CORRECTNESS OF MODEL REFINEMENT Given V∗ , our technique for model adjustment is guaranteed to choose a model M MAP ∈ M∗ by construction. If any constraint V ∗ ∈ V∗ is violated, the rejection sampling step of our technique 6 100 Comparing convergence of Different Techniques 80 70 60 50 40 30 Data Only Constraints Only Data+Constraints 20 10 0 1 2 3 4 5 Number of constraints used 6 −10 −12 −14 −16 −18 7 −20 Figure 7: Mean KLdivergence and one standard deviation for a 3 cause 3 test network on learning with data, constraints and data+constraints. Gibbs Sampling Stochastic Hill Climbing Greedy Sampling −8 Negative Log Likelihood of MAP Estimate Percentage of tests correctly predicted 90 0 1 2 3 10 10 10 10 Elapsed Time (plotted on log scale from 0 to 1500 seconds) Figure 8: Test Consistency for a 3 cause 3 test network on learning with data, constraints and data+constraints. Figure 9: Convergence rate comparison. would reject that set of parameters. To illustrate this, consider the network in Fig. 2. There are six parameters (four link reliabilities and two leak parameters). Let us fix the leak parameters and the link reliability from the first cause to each test. Now we can compute the posterior surface over the two variable parameters after discretizing each parameter in small steps and then calculating the posterior probability at each step as shown in Fig. 5. We can compare this surface with that obtained after Gibbs sampling using our technique as shown in Fig. 6. We can see that our technique recovers the posterior surface from which we can compute the MAP. We obtain the same MAP estimate with the stochastic hill climbing and greedy search algorithms. 5.3 EXPERIMENTAL RESULTS ON SYNTHETIC PROBLEMS We start by presenting our results on a 3-cause by 3-test fully-connected bipartite Bayes network. We assume that there exists some M ∗ ∈ M∗ that we want to learn given V+ . We use our technique to find M MAP . To evaluate M MAP , we first compute the constraints, V∗ for M ∗ to get the feasible region associated with the true model. Next, we sample 100 other models from this feasible region that are diagnostically equivalent. We compare these models with M MAP (after collecting 200 samples with non-informative priors for the parameters). We compute the KL-divergence of M MAP with respect to each sampled model. We expect KLdivergence to decrease as the number of constraints in V+ increases since the feasible region beMAP comes smaller. Figure 7 confirms this trend and shows that MDV+ has lower mean KL-divergence MAP MAP than MV+ , which has lower mean KL-divergence than MD . The data points in D are limited to the results of the diagnostic sessions needed to obtain V+ . As constraints increase, more data is available and so the results for the data-only approach also improve with increasing constraints. We also compare the test consistency when learning from data only, constraints only or both. Given a fixed number of constraints, we enumerate the unobserved trajectories, and then compute the highest ranked test using the learnt model and the sampled true models, for each trajectory. The test consistency is reported as a percentage, with 100% consistency indicating that the learned and true models had the same highest ranked tests on every trajectory. Figure 8 presents these percentatges for the greedy sampling technique (the results are similar for the other techniques). It again appears that learning parameters with both constraints and data is better than learning with only constraints, which is most of the times better than learning with only data. Figure 9 compares the convergence rate of each technique to find the MAP estimate. As expected, Stochastic Hill Climbing and Greedy Sampling take less time than Gibbs sampling to find parameter settings with high posterior probability. 5.4 EXPERIMENTAL RESULTS ON REAL-WORLD PROBLEMS We evaluate our technique on a real-world diagnostic network collected and reported by Agosta et al. [1], where the authors collected detailed session logs over a period of seven weeks in which the 7 KL−divergence of when computing joint over all tests 8 Figure 10: Diagnostic Bayesian network collected from user trials and pruned to retain sub-networks with at least one constraint Data Only Constraints Only Data+Constraints 7 6 5 4 3 2 1 6 8 10 12 14 16 Number of constraints used 18 20 22 Figure 11: KL divergence comparison as the number of constraints increases for the real world problem. entire diagnostic sequence was recorded. The sequences intermingle model building and querying phases. The model network structure was inferred from an expert’s sequence of positing causes and tests. Test-ranking constraints were deduced from the expert’s test query sequences once the network structure is established. The 157 sessions captured over the seven weeks resulted in a Bayes network with 115 tests, 82 root causes and 188 arcs. The network consists of several disconnected sub-networks, each identified with a symptom represented by the first test in the sequence, and all subsequent tests applied within the same subnet. There were 20 sessions from which we were able to observe trajectories with at least two tests, resulting in a total of 32 test constraints. We pruned our diagnostic network to remove the sub-networks with no constraints to get a Bayes network with 54 tests, 30 root causes, and 67 parameters divided in 7 sub-networks, as shown in Figure 10, on which we apply our model refinement technique to learn the parameters for each sub-network separately. Since we don’t have the true underlying network and the full set of constraints (more constraints could be observed in future diagnostic sessions), we treated the 32 constraints as if they were V∗ and the corresponding feasible region M∗ as if it contained models diagnostically equivalent to the unknown true model. Figure 11 reports the KL divergence between the models found by our algorithms and sampled models from M∗ as we increase the number of constraints. With such limited constraints and consequently large feasible regions, it is not surprising that the variation in KL divergence is large. Again, the MAP estimate based on both the constraints and the data has lower KL divergence than constraints only and data only. 6 CONCLUSION AND FUTURE WORK In summary, we presented an approach that can learn the parameters of a Bayes network based on constraints implied by test consistency and any data available. While several approaches exist to incorporate qualitative constraints in learning procedures, our work makes two important contributions: First, this is the first approach that exploits implicit constraints based on value of information assessments. Secondly it is the first approach that can handle non-convex constraints. We demonstrated the approach on synthetic data and on a real-world manufacturing diagnostic problem. Since data is generally sparse in diagnostics, this work makes an important advance to mitigate the model acquisition bottleneck, which has prevented the widespread application of diagnostic networks so far. In the future, it would be interesting to generalize this work to reinforcement learning in applications where data is sparse, but constraints may be inferred from expert interactions. Acknowledgments This work was supported by a grant from Intel Corporation. 8 References [1] John Mark Agosta, Omar Zia Khan, and Pascal Poupart. Evaluation results for a query-based diagnostics application. In The Fifth European Workshop on Probabilistic Graphical Models (PGM 10), Helsinki, Finland, September 13–15 2010. [2] Eric E. Altendorf, Angelo C. Restificar, and Thomas G. Dietterich. Learning from sparse data by exploiting monotonicity constraints. In Proceedings of Twenty First Conference on Uncertainty in Artificial Intelligence (UAI), Edinburgh, Scotland, July 2005. [3] Brigham S. Anderson and Andrew W. Moore. Fast information value for graphical models. In Proceedings of Nineteenth Annual Conference on Neural Information Processing Systems (NIPS), pages 51–58, Vancouver, BC, Canada, December 2005. [4] Cassio P. de Campos and Qiang Ji. Improving Bayesian network parameter learning using constraints. In International Conference in Pattern Recognition (ICPR), Tampa, FL, USA, 2008. [5] Marek J. Druzdzel and Linda C. van der Gaag. Elicitation of probabilities for belief networks: combining qualitative and quantitative information. In Proceedings of the Eleventh Annual Conference on Uncertainty in Artificial Intelligence (UAI), pages 141–148, Montreal, QC, Canada, 1995. [6] Ad J. Feelders. A new parameter learning method for Bayesian networks with qualitative influences. In Proceedings of Twenty Third International Conference on Uncertainty in Artificial Intelligence (UAI), Vancouver, BC, July 2007. [7] Mara Angeles Gil and Pedro Gil. A procedure to test the suitability of a factor for stratification in estimating diversity. Applied Mathematics and Computation, 43(3):221 – 229, 1991. [8] David Heckerman and John S. Breese. Causal independence for probability assessment and inference using bayesian networks. IEEE Systems, Man, and Cybernetics, 26(6):826–831, November 1996. [9] David Heckerman, John S. Breese, and Koos Rommelse. Decision-theoretic troubleshooting. Communications of the ACM, 38(3):49–56, 1995. [10] Ronald A. Howard. Information value theory. IEEE Transactions on Systems Science and Cybernetics, 2(1):22–26, August 1966. [11] Percy Liang, Michael I. Jordan, and Dan Klein. Learning from measurements in exponential families. In Proceedings of Twenty Sixth Annual International Conference on Machine Learning (ICML), Montreal, QC, Canada, June 2009. [12] Wenhui Liao and Qiang Ji. Learning Bayesian network parameters under incomplete data with domain knowledge. Pattern Recognition, 42:3046–3056, 2009. [13] Yi Mao and Guy Lebanon. Domain knowledge uncertainty and probabilistic parameter constraints. In Proceedings of Twenty Fifth Conference on Uncertainty in Artificial Intelligence (UAI), Montreal, QC, Canada, 2009. [14] Ryszard S. Michalski. A theory and methodology of inductive learning. Artificial Intelligence, 20:111–116, 1984. [15] Radu Stefan Niculescu, Tom M. Mitchell, and R. Bharat Rao. Bayesian network learning with parameter constraints. Journal of Machine Learning Research, 7:1357–1383, 2006. [16] Mark A. Peot and Ross D. Shachter. Learning from what you dont observe. In Proceedings of the Fourteenth Conference on Uncertainty in Artificial Intelligence (UAI), pages 439–446, Madison, WI, July 1998. [17] Michael P. Wellman. Fundamental concepts of qualitative probabilistic networks. Artificial Intelligence, 44(3):257–303, August 1990. [18] Frank Wittig and Anthony Jameson. Exploiting qualitative knowledge in the learning of conditional probabilities of Bayesian networks. In Proceedings of the Sixteenth Conference on Uncertainty in Artificial Intelligence (UAI), San Francisco, CA, July 2000. 9
3 0.45296887 33 nips-2011-An Exact Algorithm for F-Measure Maximization
Author: Krzysztof J. Dembczynski, Willem Waegeman, Weiwei Cheng, Eyke Hüllermeier
Abstract: The F-measure, originally introduced in information retrieval, is nowadays routinely used as a performance metric for problems such as binary classification, multi-label classification, and structured output prediction. Optimizing this measure remains a statistically and computationally challenging problem, since no closed-form maximizer exists. Current algorithms are approximate and typically rely on additional assumptions regarding the statistical distribution of the binary response variables. In this paper, we present an algorithm which is not only computationally efficient but also exact, regardless of the underlying distribution. The algorithm requires only a quadratic number of parameters of the joint distribution (with respect to the number of binary responses). We illustrate its practical performance by means of experimental results for multi-label classification. 1
4 0.44776636 42 nips-2011-Bayesian Bias Mitigation for Crowdsourcing
Author: Fabian L. Wauthier, Michael I. Jordan
Abstract: Biased labelers are a systemic problem in crowdsourcing, and a comprehensive toolbox for handling their responses is still being developed. A typical crowdsourcing application can be divided into three steps: data collection, data curation, and learning. At present these steps are often treated separately. We present Bayesian Bias Mitigation for Crowdsourcing (BBMC), a Bayesian model to unify all three. Most data curation methods account for the effects of labeler bias by modeling all labels as coming from a single latent truth. Our model captures the sources of bias by describing labelers as influenced by shared random effects. This approach can account for more complex bias patterns that arise in ambiguous or hard labeling tasks and allows us to merge data curation and learning into a single computation. Active learning integrates data collection with learning, but is commonly considered infeasible with Gibbs sampling inference. We propose a general approximation strategy for Markov chains to efficiently quantify the effect of a perturbation on the stationary distribution and specialize this approach to active learning. Experiments show BBMC to outperform many common heuristics. 1
5 0.44082493 147 nips-2011-Learning Patient-Specific Cancer Survival Distributions as a Sequence of Dependent Regressors
Author: Hsiu-chin Lin, Vickie Baracos, Russell Greiner, Chun-nam J. Yu
Abstract: An accurate model of patient survival time can help in the treatment and care of cancer patients. The common practice of providing survival time estimates based only on population averages for the site and stage of cancer ignores many important individual differences among patients. In this paper, we propose a local regression method for learning patient-specific survival time distribution based on patient attributes such as blood tests and clinical assessments. When tested on a cohort of more than 2000 cancer patients, our method gives survival time predictions that are much more accurate than popular survival analysis models such as the Cox and Aalen regression models. Our results also show that using patient-specific attributes can reduce the prediction error on survival time by as much as 20% when compared to using cancer site and stage only. 1
6 0.39844459 62 nips-2011-Continuous-Time Regression Models for Longitudinal Networks
7 0.38129124 137 nips-2011-Iterative Learning for Reliable Crowdsourcing Systems
8 0.3709785 116 nips-2011-Hierarchically Supervised Latent Dirichlet Allocation
9 0.35252413 240 nips-2011-Robust Multi-Class Gaussian Process Classification
10 0.34646451 6 nips-2011-A Global Structural EM Algorithm for a Model of Cancer Progression
11 0.34522837 238 nips-2011-Relative Density-Ratio Estimation for Robust Distribution Comparison
12 0.33865225 27 nips-2011-Advice Refinement in Knowledge-Based SVMs
13 0.32572654 193 nips-2011-Object Detection with Grammar Models
14 0.3223317 180 nips-2011-Multiple Instance Filtering
15 0.31480333 77 nips-2011-Efficient Learning of Generalized Linear and Single Index Models with Isotonic Regression
16 0.31393552 3 nips-2011-A Collaborative Mechanism for Crowdsourcing Prediction Problems
17 0.31263292 277 nips-2011-Submodular Multi-Label Learning
18 0.30434602 60 nips-2011-Confidence Sets for Network Structure
19 0.29816476 288 nips-2011-Thinning Measurement Models and Questionnaire Design
20 0.29441839 247 nips-2011-Semantic Labeling of 3D Point Clouds for Indoor Scenes
topicId topicWeight
[(0, 0.025), (4, 0.048), (20, 0.026), (26, 0.028), (31, 0.085), (43, 0.059), (45, 0.07), (48, 0.352), (57, 0.029), (74, 0.026), (80, 0.022), (83, 0.043), (87, 0.012), (99, 0.036)]
simIndex simValue paperId paperTitle
same-paper 1 0.79369241 232 nips-2011-Ranking annotators for crowdsourced labeling tasks
Author: Vikas C. Raykar, Shipeng Yu
Abstract: With the advent of crowdsourcing services it has become quite cheap and reasonably effective to get a dataset labeled by multiple annotators in a short amount of time. Various methods have been proposed to estimate the consensus labels by correcting for the bias of annotators with different kinds of expertise. Often we have low quality annotators or spammers–annotators who assign labels randomly (e.g., without actually looking at the instance). Spammers can make the cost of acquiring labels very expensive and can potentially degrade the quality of the consensus labels. In this paper we formalize the notion of a spammer and define a score which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a high score close to one. 1 Spammers in crowdsourced labeling tasks Annotating an unlabeled dataset is one of the bottlenecks in using supervised learning to build good predictive models. Getting a dataset labeled by experts can be expensive and time consuming. With the advent of crowdsourcing services (Amazon’s Mechanical Turk being a prime example) it has become quite easy and inexpensive to acquire labels from a large number of annotators in a short amount of time (see [8], [10], and [11] for some computer vision and natural language processing case studies). One drawback of most crowdsourcing services is that we do not have tight control over the quality of the annotators. The annotators can come from a diverse pool including genuine experts, novices, biased annotators, malicious annotators, and spammers. Hence in order to get good quality labels requestors typically get each instance labeled by multiple annotators and these multiple annotations are then consolidated either using a simple majority voting or more sophisticated methods that model and correct for the annotator biases [3, 9, 6, 7, 14] and/or task complexity [2, 13, 12]. In this paper we are interested in ranking annotators based on how spammer like each annotator is. In our context a spammer is a low quality annotator who assigns random labels (maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator). Spammers can significantly increase the cost of acquiring annotations (since they need to be paid) and at the same time decrease the accuracy of the final consensus labels. A mechanism to detect and eliminate spammers is a desirable feature for any crowdsourcing market place. For example one can give monetary bonuses to good annotators and deny payments to spammers. The main contribution of this paper is to formalize the notion of a spammer for binary, categorical, and ordinal labeling tasks. More specifically we define a scalar metric which can be used to rank the annotators—with the spammers having a score close to zero and the good annotators having a score close to one (see Figure 4). We summarize the multiple parameters corresponding to each annotator into a single score indicative of how spammer like the annotator is. While this spammer score was implicit for binary labels in earlier works [3, 9, 2, 6] the extension to categorical and ordinal labels is novel and is quite different from the accuracy computed from the confusion rate matrix. An attempt to quantify the quality of the workers based on the confusion matrix was recently made by [4] where they transformed the observed labels into posterior soft labels based on the estimated confusion 1 matrix. While we obtain somewhat similar annotator rankings, we differ from this work in that our score is directly defined in terms of the annotator parameters (see § 5 for more details). The rest of the paper is organized as follows. For ease of exposition we start with binary labels (§ 2) and later extend it to categorical (§ 3) and ordinal labels (§ 4). We first specify the annotator model used, formalize the notion of a spammer, and propose an appropriate score in terms of the annotator model parameters. We do not dwell too much on the estimation of the annotator model parameters. These parameters can either be estimated directly using known gold standard 1 or the iterative algorithms that estimate the annotator model parameters without actually knowing the gold standard [3, 9, 2, 6, 7]. In the experimental section (§ 6) we obtain rankings for the annotators using the proposed spammer scores on some publicly available data from different domains. 2 Spammer score for crowdsourced binary labels j Annotator model Let yi ∈ {0, 1} be the label assigned to the ith instance by the j th annotator, and let yi ∈ {0, 1} be the actual (unobserved) binary label. We model the accuracy of the annotator separately on the positive and the negative examples. If the true label is one, the sensitivity (true positive rate) αj for the j th annotator is defined as the probability that the annotator labels it as one. j αj := Pr[yi = 1|yi = 1]. On the other hand, if the true label is zero, the specificity (1−false positive rate) β j is defined as the probability that annotator labels it as zero. j β j := Pr[yi = 0|yi = 0]. Extensions of this basic model have been proposed to include item level difficulty [2, 13] and also to model the annotator performance based on the feature vector [14]. For simplicity we use the basic model proposed in [7] in our formulation. Based on many instances labeled by multiple annotators the maximum likelihood estimator for the annotator parameters (αj , β j ) and also the consensus ground truth (yi ) can be estimated iteratively [3, 7] via the Expectation Maximization (EM) algorithm. The EM algorithm iteratively establishes a particular gold standard (initialized via majority voting), measures the performance of the annotators given that gold standard (M-step), and refines the gold standard based on the performance measures (E-step). Who is a spammer? Intuitively, a spammer assigns labels randomly—maybe because the annotator does not understand the labeling criteria, does not look at the instances when labeling, or maybe a bot pretending to be a human annotator. More precisely an annotator is a spammer if the probability j of observed label yi being one given the true label yi is independent of the true label, i.e., j j Pr[yi = 1|yi ] = Pr[yi = 1]. This means that the annotator is assigning labels randomly by flipping a coin with bias without actually looking at the data. Equivalently (1) can be written as j j Pr[yi = 1|yi = 1] = Pr[yi = 1|yi = 0] which implies αj = 1 − β j . (1) j Pr[yi = 1] (2) Hence in the context of the annotator model defined earlier a perfect spammer is an annotator for whom αj + β j − 1 = 0. This corresponds to the diagonal line on the Receiver Operating Characteristic (ROC) plot (see Figure 1(a)) 2 . If αj + β j − 1 < 0 then the annotators lies below the diagonal line and is a malicious annotator who flips the labels. Note that a malicious annotator has discriminatory power if we can detect them and flip their labels. In fact the methods proposed in [3, 7] can automatically flip the labels for the malicious annotators. Hence we define the spammer score for an annotator as S j = (αj + β j − 1)2 (3) An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 1 One of the commonly used strategy to filter out spammers is to inject some items into the annotations with known labels. This is the strategy used by CrowdFlower (http://crowdflower.com/docs/gold). 2 Also note that (αj + β j )/2 is equal to the area shown in the plot and can be considered as a non-parametric approximation to the area under the ROC curve (AUC) based on one observed point. It is also equal to the Balanced Classification Rate (BCR). So a spammer can also be defined as having BCR or AUC equal to 0.5. 2 Equal accuracy contours (prevalence=0.5) 0.9 1 Good Annotators Biased Annotators 0.8 0.9 Sensitivity Sensitivity ( αj ) Spammers 0.5 0.4 j 0.3 j 0.6 0.5 0.4 4 0. 5 0. 3 0. 7 0. 6 0. 4 0. 5 0. 2 0. 3 0. Malicious Annotators 0.2 0.4 0.6 1−Specificity ( βj ) 0.8 0.1 1 0. 0. 2 0. 3 0. 1 0. 0.6 0.5 1 0. 2 0. 2 0. 1 0. 0.4 3 1 0. 0. 2 0. 4 0. 0.2 4 0. 5 0. 0 0 1 3 0. 0.3 0.2 Biased Annotators 4 0.7 6 0. 4 0. 0.8 7 0.3 Area = (α +β )/2 0.2 0 0 0.9 0. 8 0. 8 0. 0.7 6 0. .5 0 5 0. 0.7 [ 1−βj, αj ] 0.6 0.1 6 0. 0.8 0.7 Equal spammer score contours 1 7 0. 8 0. 9 0. Sensitivity 1 (a) Binary annotator model 0.1 1 0. 2 0. 3 0. 0.2 0.4 0.6 1−Specificity 0.8 1 1 0. 0 0 (b) Accuracy 0.2 3 0. 4 0. 0.4 0.6 1−Specificity 5 0. .6 7 0 0. 8 0. 0.8 1 (c) Spammer score Figure 1: (a) For binary labels an annotator is modeled by his/her sensitivity and specificity. A perfect spammer lies on the diagonal line on the ROC plot. (b) Contours of equal accuracy (4) and (c) equal spammer score (3). Accuracy This notion of a spammer is quite different for that of the accuracy of an annotator. An annotator with high accuracy is a good annotator but one with low accuracy is not necessarily a spammer. The accuracy is computed as 1 j Accuracyj = Pr[yi = yi ] = j Pr[yi = 1|yi = k]Pr[yi = k] = αj p + β j (1 − p), (4) k=0 where p := Pr[yi = 1] is the prevalence of the positive class. Note that accuracy depends on prevalence. Our proposed spammer score does not depend on prevalence and essentially quantifies the annotator’s inherent discriminatory power. Figure 1(b) shows the contours of equal accuracy on the ROC plot. Note that annotators below the diagonal line (malicious annotators) have low accuracy. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we can detect them and then correct for the flipping. In fact the EM algorithms [3, 7] can correctly flip the labels for the malicious annotators and hence they should not be treated as spammers. Figure 1(c) also shows the contours of equal score for our proposed score and it can be seen that the malicious annotators have a high score and only annotators along the diagonal have a low score (spammers). Log-odds Another interpretation of a spammer can be seen from the log odds. Using Bayes’ rule the posterior log-odds can be written as log j Pr[yi = 1|yi ] Pr[yi = j 0|yi ] = log j Pr[yi |yi = 1] j Pr[yi |yi = 0] + log p . 1−p Pr[y =1|y j ] p If an annotator is a spammer (i.e., (2) holds) then log Pr[yi =0|yi ] = log 1−p . Essentially the annotator j i i provides no information in updating the posterior log-odds and hence does not contribute to the estimation of the actual true label. 3 Spammer score for categorical labels Annotator model Suppose there are K ≥ 2 categories. We introduce a multinomial parameter αj = (αj , . . . , αj ) for each annotator, where c c1 cK K j αj := Pr[yi = k|yi = c] ck αj = 1. ck and k=1 αj ck The term denotes the probability that annotator j assigns class k to an instance given that the true class is c. When K = 2, αj and αj are sensitivity and specificity, respectively. 11 00 Who is a spammer? As earlier a spammer assigns labels randomly, i.e., j j Pr[yi = k|yi ] = Pr[yi = k], ∀k. 3 j j This is equivalent to Pr[yi = k|yi = c] = Pr[yi = k|yi = c ], ∀c, c , k = 1, . . . , K— which means knowing the true class label being c or c does not change the probability of the annotator’s assigned label. This indicates that the annotator j is a spammer if αj = αj k , ∀c, c , k = 1, . . . , K. ck c (5) Let Aj be the K × K confusion rate matrix with entries [Aj ]ck = αck —a spammer would have 0.50 0.50 0.50 all the rows of Aj equal, for example, Aj = 0.25 0.25 0.25 0.25 0.25 0.25 , for a three class categorical annotation problem. Essentially Aj is a rank one matrix of the form Aj = evj , for some column vector vj ∈ RK that satisfies vj e = 1, where e is column vector of ones. In the binary case we had this natural notion of spammer as an annotator for whom αj + β j − 1 was close to zero. One natural way to summarize (5) would be in terms of the distance (Frobenius norm) of the confusion matrix to the closest rank one approximation, i.e, S j := Aj − eˆj v 2 F, (6) where ˆj solves v ˆj = arg min Aj − evj v vj 2 F s.t. vj e = 1. (7) Solving (7) yields ˆj = (1/K)Aj e, which is the mean of the rows of Aj . Then from (6) we have v Sj = I− 1 ee K 2 Aj = F 1 K (αj − αj k )2 . ck c c < . . . < K. Annotator model It is conceptually easier to think of the true label to be binary, that is, yi ∈ {0, 1}. For example in mammography a lesion is either malignant (1) or benign (0) (which can be confirmed by biopsy) and the BIRADS ordinal scale is a means for the radiologist to quantify the uncertainty based on the digital mammogram. The radiologist assigns a higher value of the label if he/she thinks the true label is closer to one. As earlier we characterize each annotator by the sensitivity and the specificity, but the main difference is that we now define the sensitivity and specificity for j each ordinal label (or threshold) k ∈ {1, . . . , K}. Let αj and βk be the sensitivity and specificity k th respectively of the j annotator corresponding to the threshold k, that is, j j j αj = Pr[yi ≥ k | yi = 1] and βk = Pr[yi < k | yi = 0]. k j j Note that αj = 1, β1 = 0 and αj 1 K+1 = 0, βK+1 = 1 from this definition. Hence each annotator j j is parameterized by a set of 2(K − 1) parameters [αj , β2 , . . . , αj , βK ]. This corresponds to an 2 K empirical ROC curve for the annotator (Figure 2). 4 Who is a spammer? As earlier we define an an1 j k=1 notator j to be a spammer if Pr[yi = k|yi = 1] = 0.9 j k=2 0.8 Pr[yi = k|yi = 0] ∀k = 1, . . . , K. Note that from j 0.7 k=3 [ 1−β , α ] the annotation model we have 3 Pr[yi = k | yi = 0.6 j j j 1] = αk − αk+1 and Pr[yi = k | yi = 0] = 0.5 k=4 j j 0.4 βk+1 − βk . This implies that annotator j is a spam0.3 j j mer if αj − αj k k+1 = βk+1 − βk , ∀k = 1, . . . , K, 0.2 j j 0.1 which leads to αj + βk = αj + β1 = 1, ∀k. This 1 k j 0 0 0.2 0.4 0.6 0.8 1 means that for every k, the point (1 − βk , αj ) lies on k 1−Specificity ( β ) the diagonal line in the ROC plot shown in Figure 2. The area under the empirical ROC curve can be comFigure 2: Ordinal labels: An annotator is modK 1 puted as (see Figure 2) AUCj = 2 k=1 (αj + eled by sensitivity/specificity for each threshold. k+1 j j αj )(βk+1 − βk ), and can be used to define the folk lowing spammer score as (2AUCj − 1)2 to rank the different annotators. 3 Sensitivity ( αj ) 3 j 2 K (αj k+1 k=1 j S = + j αj )(βk+1 k − j βk ) −1 (9) With two levels this expression defaults to the binary case. An annotator is a spammer if S j is close to zero. Good annotators have S j > 0 while a perfect annotator has S j = 1. 5 Previous work Recently Ipeirotis et.al. [4] proposed a score for categorical labels based on the expected cost of the posterior label. In this section we briefly describe their approach and compare it with our proposed score. For each instance labeled by the annotator they first compute the posterior (soft) label j j Pr[yi = c|yi ] for c = 1, . . . , K, where yi is the label assigned to the ith instance by the j th annotator and yi is the true unknown label. The posterior label is computed via Bayes’ rule as j j j Pr[yi = c|yi ] ∝ Pr[yi |yi = c]Pr[yi = c] = (αj )δ(yi ,k) pc , where pc = Pr[yi = c] is the prevack lence of class c. The score for a spammer is based on the intuition that the posterior label vector j j (Pr[yi = 1|yi ], . . . , Pr[yi = K|yi ]) for a good annotator will have all the probability mass concentrated on single class. For example for a three class problem (with equal prevalence), a posterior label vector of (1, 0, 0) (certain that the class is one) comes from a good annotator while a (1/3, 1/3, 1/3) (complete uncertainty about the class label) comes from spammer. Based on this they define the following score for each annotator 1 Score = N N K K j j costck Pr[yi = k|yi ]Pr[yi = c|yi ] j i=1 . (10) c=1 k=1 where costck is the misclassification cost when an instance of class c is classified as k. Essentially this is capturing some sort of uncertainty of the posterior label averaged over all the instances. Perfect workers have a score Scorej = 0 while spammers will have high score. An entropic version of this score based on similar ideas has also been recently proposed in [5]. Our proposed spammer score differs from this approach in the following aspects: (1) Implicit in the score defined above (10) j is the assumption that an annotator is a spammer when Pr[yi = c|yi ] = Pr[yi = c], i.e., the estimated posterior labels are simply based on the prevalence and do not depend on the observed labels. By j j Bayes’ rule this is equivalent to Pr[yi |yi = c] = Pr[yi ] which is what we have used to define our spammer score. (2) While both notions of a spammer are equivalent, the approach of [4] first computes the posterior labels based on the observed data, the class prevalence and the annotator j j j j This can be seen as follows: Pr[yi = k | yi = 1] = Pr[(yi ≥ k) AND (yi < k + 1) | yi = 1] = Pr[yi ≥ j j j j k | yi = 1] + Pr[yi < k + 1 | yi = 1] − Pr[(yi ≥ k) OR (yi < k + 1) | yi = 1] = Pr[yi ≥ k | yi = j j j 1] − Pr[yi ≥ k + 1 | yi = 1] = αj − αj . Here we used the fact that Pr[(yi ≥ k) OR (yi < k + 1)] = 1. k k+1 3 5 simulated | 500 instances | 30 annotators simulated | 500 instances | 30 annotators 1 12 0.8 Spammer Score 18 0.6 0.5 22 24 23 25 0.3 29 20 0.2 0.4 0.2 30 16 14 0.1 26 21 27 28 19 0 0 13 0 0.2 0.4 0.6 1−Specificity 0.8 1 500 500 500 500 500 500 500 500 500 500 0.4 0.6 500 500 500 1 0.7 500 500 500 500 500 500 500 500 500 500 500 500 500 500 3 1 500 500 500 2 8 510 7 17 4 9 27 8 30 6 3 28 7 10 2 23 22 26 24 5 1 21 29 25 14 12 17 11 18 20 19 15 16 13 4 0.8 Sensitivity 6 9 0.9 15 11 Annotator (a) Simulation setup (b) Annotator ranking Annotator rank (median) via accuracy simulated | 500 instances | 30 annotators Annotator rank (median) via Ipeirotis et.al.[4] simulated | 500 instances | 30 annotators 27 30 30 28 23 22 26 25 24 21 29 25 20 14 17 18 15 20 16 15 11 13 12 19 1 10 5 10 2 7 6 3 5 8 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (c) Comparison with accuracy 30 18 20 16 15 19 13 12 25 11 14 17 25 20 29 21 51 15 26 22 23 24 10 2 7 10 28 6 3 8 30 5 27 9 4 0 0 5 10 15 20 25 Annotator rank (median) via spammer score 30 (d) Comparison with Ipeirotis et. al. [4] Figure 3: (a) The simulation setup consisting of 10 good annotators (annotators 1 to 10), 10 spammers (11 to 20), and 10 malicious annotators (21 to 30). (b) The ranking of annotators obtained using the proposed spammer score. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. (c) and (d) Comparison of the median rank obtained via the spammer score with the rank obtained using (c) accuracy and (d) the method proposed by Ipeirotis et. al. [4]. parameters and then computes the expected cost. Our proposed spammer score does not depend on the prevalence of the class. Our score is also directly defined only in terms of the annotator confusion matrix and does not need the observed labels. (3) For the score defined in (10) while perfect annotators have a score of 0 it is not clear what should be a good baseline for a spammer. The authors suggest to compute the baseline by assuming that a worker assigns as label the class with maximum prevalence. Our proposed score has a natural scale with a perfect annotator having a score of 1 and a spammer having a score of 0. (4) However one advantage of the approach in [4] is that they can directly incorporate varied misclassification costs. 6 Experiments Ranking annotators based on the confidence interval As mentioned earlier the annotator model parameters can be estimated using the iterative EM algorithms [3, 7] and these estimated annotator parameters can then be used to compute the spammer score. The spammer score can then be used to rank the annotators. However one commonly observed phenomenon when working with crowdsourced data is that we have a lot of annotators who label only a very few instances. As a result the annotator parameters cannot be reliably estimated for these annotators. In order to factor this uncertainty in the estimation of the model parameters we compute the spammer score for 100 bootstrap replications. Based on this we compute the 95% confidence intervals (CI) for the spammer score for each annotator. We rank the annotators based on the lower limit of the 95% CI. The CIs are wider 6 Table 1: Datasets N is the number of instances. M is the number of annotators. M ∗ is the mean/median number of annotators per instance. N ∗ is the mean/median number of instances labeled by each annotator. Dataset Type N M M∗ N∗ bluebird binary 108 39 39/39 108/108 temp binary 462 76 10/10 61/16 Brief Description wsd categorical/3 177 34 10/10 52/20 sentiment categorical/3 1660 33 6/6 291/175 30 100 10 38 10/10 10/10 bird identification [12] The annotator had to identify whether there was an Indigo Bunting or Blue Grosbeak in the image. event annotation [10] Given a dialogue and a pair of verbs annotators need to label whether the event described by the first verb occurs before or after the second. 30/30 26/20 30 30 30 30 30 30 3 30 30 1 30 20 20 20 20 20 77 117 20 60 Spammer Score 0.4 10 7 9 8 6 5 0 2 0.2 13 31 10 23 29 1 2 4 6 8 9 14 15 17 22 32 5 18 16 19 11 12 20 21 24 25 26 27 28 30 33 34 7 3 0 0.6 20 20 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 108 20 20 20 20 17 17 40 20 20 100 Spammer Score 0.4 108 108 108 108 108 108 108 108 108 108 108 108 108 0.8 0.6 0.2 17 8 27 30 25 35 1 12 32 37 38 16 22 9 29 15 20 19 5 39 3 21 23 14 2 10 24 7 33 13 36 31 4 34 28 18 11 6 26 0.2 30 77 77 4 108 108 108 108 0.4 0 wosi | 30 instances | 10 annotators 1 0.8 108 108 0.6 108 108 108 Spammer Score 0.8 wsd | 177 instances | 34 annotators 1 80 177 157 177 157 bluebird | 108 instances | 39 annotators 1 word similarity [10] Numeric judgements of word similarity. affect recognition [10] Each annotator is presented with a short headline and asked to rate it on a scale [-100,100] to denote the overall positive or negative valence. 40 40 20 ordinal/[0 10] ordinal[-100 100] 20 20 20 wosi valence word sense disambiguation [10] The labeler is given a paragraph of text containing the word ”president” and asked to label one of the three appropriate senses. irish economic sentiment analysis [1] Articles from three Irish online news sources were annotated by volunteer users as positive, negative, or irrelevant. 20 20 20 20 20 60 20 20 20 40 40 100 0.4 Annotator Annotator 0 1 26 10 18 28 15 5 36 23 12 8 32 31 38 13 17 27 11 2 35 24 19 9 6 30 33 37 14 29 4 3 20 34 22 25 7 16 21 40 20 0.2 26 2 6 11 5 14 3 20 9 22 31 10 12 18 8 13 30 4 1 29 19 17 27 28 21 15 25 23 7 33 16 24 32 10 132 10 360 10 0 13 18 52 75 33 32 12 74 31 51 41 55 7 14 70 42 58 65 43 1 10 47 61 73 25 37 76 67 24 46 54 48 39 56 15 62 68 44 53 64 40 9 28 6 2 57 3 4 5 8 11 16 17 19 20 21 22 23 26 27 29 30 34 35 36 38 45 49 50 59 60 63 66 69 71 72 442 462 452 10 10 0.6 238 171 75 654 20 0.2 0.2 0 12 77 67 374 249 229 453 346 428 0.4 Spammer Score 43 175 119 541 525 437 0.8 917 104 284 0.4 0.6 1211 1099 10 Spammer Score 0.8 572 30 52 402 60 0.6 30 Spammer Score 0.8 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 60 40 20 15 7 7 11 12 35 29 1 87 10 10 10 10 10 10 10 12 1 30 10 10 10 10 10 10 10 10 10 10 10 10 10 valence | 100 instances | 38 annotators 20 22 10 10 10 10 sentiment | 1660 instances | 33 annotators 10 10 30 20 10 Annotator temp | 462 instances | 76 annotators 1 Annotator 10 50 10 10 40 10 70 350 80 40 100 192 190 40 32 60 70 20 20 40 80 20 50 50 50 30 Annotator Annotator Figure 4: Annotator Rankings The rankings obtained for the datasets in Table 1. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence intervals (CI) are shown—obtained from 100 bootstrap replications. The annotators are ranked based on the lower limit of the 95% CI. The number at the top of the CI bar shows the number of instances annotated by that annotator. Note that the CIs are wider when the annotator labels only a few instances. when the annotator labels only a few instances. For a crowdsourced labeling task the annotator has to be good and also label a reasonable number of instances in order to be reliably identified. Simulated data We first illustrate our proposed spammer score on simulated binary data (with equal prevalence for both classes) consisting of 500 instances labeled by 30 annotators of varying sensitivity and specificity (see Figure 3(a) for the simulation setup). Of the 30 annotators we have 10 good annotators (annotators 1 to 10 who lie above the diagonal in Figure 3(a)), 10 spammers (annotators 11 to 20 who lie around the diagonal), and 10 malicious annotators (annotators 21 to 30 who lie below the diagonal). Figure 3(b) plots the ranking of annotators obtained using the proposed spammer score with the annotator model parameters estimated via the EM algorithm [3, 7]. The spammer score ranges from 0 to 1, the lower the score, the more spammy the annotator. The mean spammer score and the 95% confidence interval (CI) obtained via bootstrapping are shown. The annotators are ranked based on the lower limit of the 95% CI. As can be seen all the spammers (annotators 11 to 20) have a low spammer score and appear at the bottom of the list. The malicious annotators have higher score than the spammers since we can correct for their flipping. The malicious annotators are good annotators but they flip their labels and as such are not spammers if we detect that they are malicious. Figure 3(c) compares the (median) rank obtained via the spammer score with the (median) rank obtained using accuracy as the score to rank the annotators. While the good annotators are ranked high by both methods the accuracy score gives a low rank to the malicious annotators. Accuracy does not capture the notion of a spammer. Figure 3(d) compares the ranking with the method proposed by Ipeirotis et. al. [4] which gives almost similar rankings as our proposed score. 7 21 23 10 6 35 4 34 1126 18 147 30 3 31 13 2436 33 25 5 2 20 15 39 19 15 20 28 22 299 12 37 16 38 10 32 1 5 27 25 35 30 8 17 0 0 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 40 bluebird | 108 instances | 39 annotators 40 1 6 34 112618 4 31 1013 7 30 2 28 21 5 20 15 39 19 20 15 22 37 16 299 12 38 10 5 8 17 0 0 27 25 35 30 0.6 0.5 35 32 2 0.4 36 11 13 31 24 10 33 28 21 26 18 0 0 40 34 15 19 39 0.1 (a) 22 37 20 38 29 9 0.2 5 10 15 20 25 30 35 Annotator rank (median) via spammer score 6 4 16 0.3 32 1 7 30 25 1 3 14 3 27 5 0.7 24 33 14 36 23 25 12 8 0.9 17 0.8 35 Sensitivity Annotator rank (median) via accuracy bluebird | 108 instances | 39 annotators Annotator rank (median) via Ipeirotis et.al.[4] bluebird | 108 instances | 39 annotators 40 23 0.2 0.4 0.6 1−Specificity (b) 0.8 1 (c) Figure 5: Comparison of the rank obtained via the spammer score with the rank obtained using (a) accuracy and (b) the method proposed by Ipeirotis et. al. [4] for the bluebird binary dataset. (c) The annotator model parameters as estimated by the EM algorithm [3, 7]. 19 25 12 18 7 3 14 20 32 5 8 1 16 20 9 21 15 34 10 31 29 17 28 22 26 2315 5 2 0 0 4 6 13 10 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 30 25 16 19 7 25 8 9 27 14 3 28 17 18 32 5 10 4 2 10 6 1529 31 23 22 21 15 0 0 33 30 11 1 20 5 sentiment | 1660 instances | 33 annotators 24 35 12 20 24 34 26 13 5 10 15 20 25 30 Annotator rank (median) via spammer score 35 33 7 30 15 17 25 28 2719 2223 20 8 1 4 1812 15 13 10 20 32 30 10 3 29 9 31 16 5 6 2 5 14 11 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score 25 21 Annotator rank (median) via Ipeirotis et.al.[4] 25 24 27 Annotator rank (median) via accuracy Annotator rank (median) via accuracy 30 sentiment | 1660 instances | 33 annotators wsd | 177 instances | 34 annotators 33 30 11 Annotator rank (median) via Ipeirotis et.al.[4] wsd | 177 instances | 34 annotators 35 7 30 15 19 17 27 25 21 25 8 12 4 18 20 24 15 20 33 10 3 13 9 28 1 29 23 10 1632 11 14 5 6 2 5 31 30 22 26 0 0 5 10 15 20 25 30 Annotator rank (median) via spammer score Figure 6: Comparison of the median rank obtained via the spammer score with the rank obtained using accuracy and he method proposed by Ipeirotis et. al. [4] for the two categorial datasets in Table 1. Mechanical Turk data We report results on some publicly available linguistic and image annotation data collected using the Amazon’s Mechanical Turk (AMT) and other sources. Table 1 summarizes the datasets. Figure 4 plots the spammer scores and rankings obtained. The mean and the 95% CI obtained via bootstrapping are also shown. The number at the top of the CI bar shows the number of instances annotated by that annotator. The rankings are based on the lower limit of the 95% CI which factors the number of instances labeled by the annotator into the ranking. An annotator who labels only a few instances will have very wide CI. Some annotators who label only a few instances may have a high mean spammer score but the CI will be wide and hence ranked lower. Ideally we would like to have annotators with a high score and at the same time label a lot of instances so that we can reliablly identify them. The authors [1] for the sentiment dataset shared with us some of the qualitative observations regarding the annotators and they somewhat agree with our rankings. For example the authors made the following comments about Annotator 7 ”Quirky annotator - had a lot of debate about what was the meaning of the annotation question. I’d say he changed his labeling strategy at least once during the process”. Our proposed score gave a low rank to this annotator. Comparison with other approaches Figure 5 and 6 compares the proposed ranking with the rank obtained using accuracy and the method proposed by Ipeirotis et. al. [4] for some binary and categorical datasets in Table 1. Our proposed ranking is somewhat similar to that obtained by Ipeirotis et. al. [4] but accuracy does not quite capture the notion of spammer. For example for the bluebird dataset for annotator 21 (see Figure 5(a)) accuracy ranks it at the bottom of the list while the proposed score puts is in the middle of the list. From the estimated model parameters it can be seen that annotator 21 actually flips the labels (below the diagonal in Figure 5(c)) but is a good annotator. 7 Conclusions We proposed a score to rank annotators for crowdsourced binary, categorical, and ordinal labeling tasks. The obtained rankings and the scores can be used to allocate monetary bonuses to be paid to different annotators and also to eliminate spammers from further labeling tasks. A mechanism to rank annotators should be desirable feature of any crowdsourcing service. The proposed score should also be useful to specify the prior for Bayesian approaches to consolidate annotations. 8 References [1] A. Brew, D. Greene, and P. Cunningham. Using crowdsourcing and active learning to track sentiment in online media. In Proceedings of the 6th Conference on Prestigious Applications of Intelligent Systems (PAIS’10), 2010. [2] B. Carpenter. Multilevel bayesian models of categorical data annotation. Technical Report available at http://lingpipe-blog.com/lingpipe-white-papers/, 2008. [3] A. P. Dawid and A. M. Skene. Maximum likeihood estimation of observer error-rates using the EM algorithm. Applied Statistics, 28(1):20–28, 1979. [4] P. G. Ipeirotis, F. Provost, and J. Wang. Quality management on Amazon Mechanical Turk. In Proceedings of the ACM SIGKDD Workshop on Human Computation (HCOMP’10), pages 64–67, 2010. [5] V. C. Raykar and S. Yu. An entropic score to rank annotators for crowdsourced labelling tasks. In Proceedings of the Third National Conference on Computer Vision, Pattern Recognition, Image Processing and Graphics (NCVPRIPG), 2011. [6] V. C. Raykar, S. Yu, L .H. Zhao, A. Jerebko, C. Florin, G. H. Valadez, L. Bogoni, and L. Moy. Supervised learning from multiple experts: Whom to trust when everyone lies a bit. In Proceedings of the 26th International Conference on Machine Learning (ICML 2009), pages 889– 896, 2009. [7] V. C. Raykar, S. Yu, L. H. Zhao, G. H. Valadez, C. Florin, L. Bogoni, and L. Moy. Learning from crowds. Journal of Machine Learning Research, 11:1297–1322, April 2010. [8] V. S. Sheng, F. Provost, and P. G. Ipeirotis. Get another label? Improving data quality and data mining using multiple, noisy labelers. In Proceedings of the 14th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pages 614–622, 2008. [9] P. Smyth, U. Fayyad, M. Burl, P. Perona, and P. Baldi. Inferring ground truth from subjective labelling of venus images. In Advances in Neural Information Processing Systems 7, pages 1085–1092. 1995. [10] R. Snow, B. O’Connor, D. Jurafsky, and A. Y. Ng. Cheap and Fast—but is it good? Evaluating Non-Expert Annotations for Natural Language Tasks. In Proceedings of the Conference on Empirical Methods in Natural Language Processing (EMNLP ’08), pages 254–263, 2008. [11] A. Sorokin and D. Forsyth. Utility data annotation with Amazon Mechanical Turk. In Proceedings of the First IEEE Workshop on Internet Vision at CVPR 08, pages 1–8, 2008. [12] P. Welinder, S. Branson, S. Belongie, and P. Perona. The multidimensional wisdom of crowds. In Advances in Neural Information Processing Systems 23, pages 2424–2432. 2010. [13] J. Whitehill, P. Ruvolo, T. Wu, J. Bergsma, and J. Movellan. Whose vote should count more: Optimal integration of labels from labelers of unknown expertise. In Advances in Neural Information Processing Systems 22, pages 2035–2043. 2009. [14] Y. Yan, R. Rosales, G. Fung, M. Schmidt, G. Hermosillo, L. Bogoni, L. Moy, and J. Dy. Modeling annotator expertise: Learning when everybody knows a bit of something. In Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics (AISTATS 2010), pages 932–939, 2010. 9
2 0.6411069 236 nips-2011-Regularized Laplacian Estimation and Fast Eigenvector Approximation
Author: Patrick O. Perry, Michael W. Mahoney
Abstract: Recently, Mahoney and Orecchia demonstrated that popular diffusion-based procedures to compute a quick approximation to the first nontrivial eigenvector of a data graph Laplacian exactly solve certain regularized Semi-Definite Programs (SDPs). In this paper, we extend that result by providing a statistical interpretation of their approximation procedure. Our interpretation will be analogous to the manner in which 2 -regularized or 1 -regularized 2 -regression (often called Ridge regression and Lasso regression, respectively) can be interpreted in terms of a Gaussian prior or a Laplace prior, respectively, on the coefficient vector of the regression problem. Our framework will imply that the solutions to the MahoneyOrecchia regularized SDP can be interpreted as regularized estimates of the pseudoinverse of the graph Laplacian. Conversely, it will imply that the solution to this regularized estimation problem can be computed very quickly by running, e.g., the fast diffusion-based PageRank procedure for computing an approximation to the first nontrivial eigenvector of the graph Laplacian. Empirical results are also provided to illustrate the manner in which approximate eigenvector computation implicitly performs statistical regularization, relative to running the corresponding exact algorithm. 1
3 0.58808607 114 nips-2011-Hierarchical Multitask Structured Output Learning for Large-scale Sequence Segmentation
Author: Nico Goernitz, Christian Widmer, Georg Zeller, Andre Kahles, Gunnar Rätsch, Sören Sonnenburg
Abstract: We present a novel regularization-based Multitask Learning (MTL) formulation for Structured Output (SO) prediction for the case of hierarchical task relations. Structured output prediction often leads to difficult inference problems and hence requires large amounts of training data to obtain accurate models. We propose to use MTL to exploit additional information from related learning tasks by means of hierarchical regularization. Training SO models on the combined set of examples from multiple tasks can easily become infeasible for real world applications. To be able to solve the optimization problems underlying multitask structured output learning, we propose an efficient algorithm based on bundle-methods. We demonstrate the performance of our approach in applications from the domain of computational biology addressing the key problem of gene finding. We show that 1) our proposed solver achieves much faster convergence than previous methods and 2) that the Hierarchical SO-MTL approach outperforms considered non-MTL methods. 1
4 0.58411676 204 nips-2011-Online Learning: Stochastic, Constrained, and Smoothed Adversaries
Author: Alexander Rakhlin, Karthik Sridharan, Ambuj Tewari
Abstract: Learning theory has largely focused on two main learning scenarios: the classical statistical setting where instances are drawn i.i.d. from a fixed distribution, and the adversarial scenario wherein, at every time step, an adversarially chosen instance is revealed to the player. It can be argued that in the real world neither of these assumptions is reasonable. We define the minimax value of a game where the adversary is restricted in his moves, capturing stochastic and non-stochastic assumptions on data. Building on the sequential symmetrization approach, we define a notion of distribution-dependent Rademacher complexity for the spectrum of problems ranging from i.i.d. to worst-case. The bounds let us immediately deduce variation-type bounds. We study a smoothed online learning scenario and show that exponentially small amount of noise can make function classes with infinite Littlestone dimension learnable. 1
5 0.57599431 198 nips-2011-On U-processes and clustering performance
Author: Stéphan J. Clémençcon
Abstract: Many clustering techniques aim at optimizing empirical criteria that are of the form of a U -statistic of degree two. Given a measure of dissimilarity between pairs of observations, the goal is to minimize the within cluster point scatter over a class of partitions of the feature space. It is the purpose of this paper to define a general statistical framework, relying on the theory of U -processes, for studying the performance of such clustering methods. In this setup, under adequate assumptions on the complexity of the subsets forming the partition candidates, the √ excess of clustering risk is proved to be of the order OP (1/ n). Based on recent results related to the tail behavior of degenerate U -processes, it is also shown how to establish tighter rate bounds. Model selection issues, related to the number of clusters forming the data partition in particular, are also considered. 1
7 0.42467549 106 nips-2011-Generalizing from Several Related Classification Tasks to a New Unlabeled Sample
8 0.40918317 80 nips-2011-Efficient Online Learning via Randomized Rounding
9 0.39498496 75 nips-2011-Dynamical segmentation of single trials from population neural data
10 0.38914293 42 nips-2011-Bayesian Bias Mitigation for Crowdsourcing
11 0.3863875 229 nips-2011-Query-Aware MCMC
12 0.38535127 206 nips-2011-Optimal Reinforcement Learning for Gaussian Systems
13 0.38483587 258 nips-2011-Sparse Bayesian Multi-Task Learning
14 0.38467336 301 nips-2011-Variational Gaussian Process Dynamical Systems
15 0.38463283 180 nips-2011-Multiple Instance Filtering
16 0.38454056 37 nips-2011-Analytical Results for the Error in Filtering of Gaussian Processes
17 0.38351339 102 nips-2011-Generalised Coupled Tensor Factorisation
18 0.38342971 286 nips-2011-The Local Rademacher Complexity of Lp-Norm Multiple Kernel Learning
19 0.38311484 135 nips-2011-Information Rates and Optimal Decoding in Large Neural Populations
20 0.38275626 273 nips-2011-Structural equations and divisive normalization for energy-dependent component analysis