Chapter 32 Search and classification performances

32.1 TBA How much finished

10%

32.2 Introduction

The preceding chapter described the radiological search model (RSM) for FROC data. This chapter describes predictions of the RSM and how they compare with evidence. The starting point is the inferred ROC curve. While mathematically rather complicated, the results are important because they are needed to derive the ROC-likelihood function, which is used to estimate RSM parameters from ROC data in TBA Chapter 19. The preceding sentence should lead the inquisitive reader to the question: since the ROC paradigm ignores search, how is it possible to derive parameters of a model of search from the ROC curve? The answer is that the shape of the ROC curve contains information about the RSM parameters. It is fundamentally different from predictions of all conventional ROC models: binormal (Dorfman and Alf 1969), contaminated binormal model (Dorfman and Berbaum 2000), bigamma (Dorfman et al. 1997) and proper ROC (Metz and Pan 1999), namely it has a constrained end-point property, while all other models predict that the end-point, namely the uppermost non-trivial point on the ROC, reached at infinitely low reporting threshold, is (1,1), while the RSM predicts it does not reach (1,1). The nature of search is such that the limiting end-point is constrained to be below and to the left of (1,1). This key difference, allows one to estimate search parameters from ROC data. Next, the RSM is used to predict FROC and AFROC curves. Two following sections show how search performance and lesion-classification performance can be quantified from the location of the ROC end-point. Search performance is the ability to find lesions while avoiding finding non-lesions, and lesion-classification performance is the ability, having found a suspicious region, to correctly classify it; if classified as a NL it would not be marked (in the mind of the observer every mark is a potential LL, albeit at different confidence levels). Note that lesion-classification is different from classification between diseased and non-diseased cases, which is measured by the ROC-AUC. Based on the ROC/FROC/AFROC curve predictions of the RSM, a comparison is presented between area measures that can be calculated from FROC data, and this leads to an important conclusion, namely the FROC curve is a poor descriptor of search performance and that the AFROC/wAFROC are preferred. This will come as a surprise (shock?) to most researchers somewhat familiar with this field, since the overwhelming majority of users of FROC methods, particularly in CAD, have relied on the FROC curve. Finally, evidence for the validity of the RSM is presented.

32.3 Quantifying search performance #rsm-search-search-performance}

Fig. 6 here Fig. 17.6: This figure shows a typical population ROC curve labeled (a) predicted by ROC models that do not account for search performance. Search performance is defined as the ability to find lesions while avoiding non-lesions. The end-point of such a curve is at (1,1), denoted by the filled circle. By adopting a sufficiently low reporting threshold the observer can continuously move the operating point from (0,0) to (1,1). The curve labeled (b) is a typical RSM-predicted ROC curve. The end-point is downward and leftward shifted relative to (1,1), as indicated by the filled square. The observer cannot move the operating point continuously all the way from (0,0) to (1,1) because a constrained fraction of images contain no marks. The fractions of unmarked non-diseased and diseased cases determine the location of the end-point, respectively. The observer can move the operating point continuously from the origin to the end-point and no further. The location of the end-point is a measure of search performance. Higher search performance is characterized by the end-point moving upwards and to the left, ideally to (0,1) which corresponds to perfect search performance. The perpendicular distance from the end-point to the chance diagonal (c) multiplied by √2, i.e., , is defined as a measure of search performance. Lesion-classification performance is defined as the implied AUC of two unit variance normal distributions separated by the parameter of the search model. It measures the ability, having found a suspicious region, to correctly classify it as a true lesion. The code for this plot is in file mainQuantifySearchPerformance.R.

In Fig. 17.6, the line labeled (a) is a conventional model ROC curve ending at (1,1), the filled circle, while (b) shows a typical search model ROC curve ending at a point below and to the left of (1,1), the filled square. The location of the end-point of the RSM-predicted curve determines the search performance of the observer. The square root of two times the perpendicular distance (the subscript s is for search) from the end-point to the chance diagonal in Fig. 17.6, the line labeled (c), is defined as search-performance, denoted S. For example, if and then the end-point is (0,1) and S = 1. This observer has perfect search performance since no NLs are found and all lesions were found; the perpendicular distance from (0,1) to the chance diagonal is 1/√2, which multiplied by √2 yields unity. Search performance ranges from 0 to 1. Using geometry, Eqn. (17.1) and Eqn. (17.2), it follows that:

. . (17.37)

Therefore, search performance is given by:

. . (17.38)

The second form in Eqn. (17.38) shows S in terms of the physical (i.e., primed) parameters; it shows that search performance is the product of two terms: the probability of finding lesions times the probability of avoiding finding any non-lesions. This puts into a mathematical form the qualitative definition of search performance as the ability to find lesions while avoiding finding non-lesions. Since at least one parameter is needed to describe each of these probabilities, quantifying search requires at least two parameters.

Applying this definition to the case of the rational observer who does not generates any marks, one sees that the observer’s search performance is zero . This emphasizes the point that not generating NLs is not enough; one must also be able to find lesions. It is also consistent with the fact that the origin lies on the positive diagonal of the ROC, implying zero perpendicular distance between them, i.e., .

32.4 Quantifying classification performance

To avoid misunderstanding, I emphasize that lesion-classification performance is being used in a different sense from that used in ROC methodology, where classification is between diseased and non-diseased cases, not between diseased and non-diseased regions, i.e., latent NLs and latent LLs, as in the current context.

Having found a suspicious region, how good is the observer at correctly classifying true lesions and non-lesions? Lesion-classification performance C is determined by the parameter, and is defined by the implied AUC of unit variance normal distributions separated by .

. (17.39)

It ranges from 0.5 to 1. Only one parameter is needed for this, so one needs three parameters to quantify search and lesion-classification performance.

32.4.1 Lesion-classification performance and the 2AFC LKE task

It should be obvious that lesion-classification performance is similar to what is commonly measured in model-observer research using the location-known-exactly (LKE) paradigm. In this paradigm, one uses 2AFC methods as in Fig. 4.3, but one could use the ratings method as long as the lesion is cued (i.e., pointed to). On diseased cases, the lesion is cued, but to control for false positives, one must also cue a similar region on non-diseased cases, as in Fig. 4.3. In that figure, the lesion, present in one of the two images, is always in the center of one of the two fields. Sometimes cross hairs are used to indicate where the observer should be looking. The probability of a correct choice in the 2AFC task is , i.e., AUC conditioned on the (possible) position of the lesion being cued. Since the lesion is cued, search performance of the observer is irrelevant, and one expects . The reason for the inequality is that on a non-diseased case, the location being cued, in all likelihood, does not correspond to a latent NL found by the observer’s search mechanism. Latent NLs are more suspicious for disease than other locations in the case. measures the separation parameter between latent NLs and LLs. The separation parameter between latent LLs and a researcher chosen location is likely to be larger. This is because latent NLs are more suspicious for disease than a researcher chosen location. It is known that performance under this condition exceeds that in a free-search 2AFC or ROC study, denoted AUC, where the lesion is not cued and it could be anywhere. This should be obvious – pointing to the possible location of the lesion takes out the need for searching the rest of the image, which introduces the possibility of not finding the lesion and / or finding non-lesions. One expects the following ordering: . is expected to be the least, as there is uncertainty about possible lesion location. is expected to be next in order, as now uncertainty has been reduced, and the observer’s task is to pick between two cued locations, one a latent NL and the other a latent LL. is expected to be highest, as now the observer’s task is to pick between two cued locations, one a latent LL and the other a researcher chosen location, most likely not a latent NL. Data supporting the expected inequality is presented in §19.5.4.6.

32.4.2 Significance of measuring search and lesion-classification performance

The ability to quantify search and lesion-classification performance from a single paradigm (ROC) study is highly significant, going well-beyond modeling the ROC curve. ROC-AUC measures how well an observer is able to separate two groups of patients, a group of diseased patients from a group of non-diseased patients. While important, it does not inform us about how the observer goes about doing this and what is limiting performance (an exception the CBM model which yields information about how good the observer is at finding lesions but does not account for the ability of the observer to avoid NLs on non-diseased cases). In contrast, the search and lesion-classification measures described above can be used as a “diagnostic aid” in determining what is limiting performance. If search performance is poor, it indicates that the observer needs to be trained on many non-diseased cases, and learn the variants of non-diseased anatomy and learn not to confuse them for lesions. On the other hand, if lesion-classification performance is poor, then one needs to train the observer using images where the location of a possible lesion is cued, and the observer’s task is to determine if the cued location is a real lesion. The classic example here is breast CAD, where the designer level ROC curve goes almost all the way to (1,1) implying poor search performance, while lesion-classification performance could actually be quite good, because CAD has access to the pixel values and the ability to apply complex algorithms to properly classify lesions as benign or malignant.

Of course, before one can realize these benefits, one needs a way of estimating the end-point shown in Fig. 17.6 plot (b). The observer will generally not oblige by reporting every suspicious region. RSM based curve fitting is needed to estimate the end-point’s location, Chapter 19.

32.5 Discussion / Summary

This chapter has detailed ROC, FROC and AFROC curves predicted by the radiological search model (RSM). All RSM-predicted curves share the constrained end-point property that is qualitatively different from previous ROC models. In my experience, it is a property that most researchers in this field have difficulty accepting. There is too much history going back to the early 1940s, of the ROC curve extending from (0,0) to (1,1) that one has to let go of, and this can be difficult.

I am not aware of any direct evidence that radiologists can move the operating point continuously in the range (0,0) to (1,1) in search tasks, so the existence of such an ROC is tantamount to an assumption. Algorithmic observers that do not involve the element of search can extend continuously to (1,1). An example of an algorithmic observer not involving search is a diagnostic test that rates the results of a laboratory measurement, e.g., the A1C measure of blood glucose for presence of a disease. If A1C ≥ 6.5% the patient is diagnosed as diabetic. By moving the threshold from infinity to –infinity, and assuming a large population of patients, one can trace out the entire ROC curve from the origin to (1,1). This is because every patient yields an A1C value. Now imagine that some finite fraction of the test results are “lost in the mail”; then the ROC curve, calculated over all patients, would have the constrained end-point property, albeit due to an unreasonable cause.

The situation in medical imaging involving search tasks is qualitatively different. Not every case yields a decision variable. There is a reasonable cause for this – to render a decision variable sample the radiologist must find something suspicious to report, and if none is found, there is no decision variable to report. The ROC curve calculated over all patients would exhibit the constrained end-point property, even in the limit of an infinite number of patients. If calculated over only those patients that yielded at least one mark, the ROC curve would extend from (0,0) to (1,1) but then one would be ignoring the cases with no marks, which represent valuable information: unmarked non-diseased cases represent perfect decisions and unmarked diseased cases represent worst-case decisions.

RSM-predicted ROC, FROC and AFROC curves were derived (wAFROC is implemented in the Rjafroc). These were used to demonstrate that the FROC is a poor descriptor of performance. Since almost all work to date, including some by me 47,48, has used FROC curves to measure performance, this is going to be difficulty for some to accept. The examples in Fig. 17.6 (A- F) and Fig. 17.7 (A-B) should convince one that the FROC curve is indeed a poor measure of performance. The only situation where one can safely use the FROC curve is if the two modalities produce curves extending over the same NLF range. This can happen with two variants of a CAD algorithm, but rarely with radiologist observers.

A unique feature is that the RSM provides measures of search and lesion-classification performance. It bears repeating that search performance is the ability to find lesions while avoiding finding non-lesions. Search performance can be determined from the position of the ROC end-point (which in turn is determined by RSM-based fitting of ROC data, Chapter 19). The perpendicular distance between the end-point and the chance diagonal is, apart from a factor of 1.414, a measure of search performance. All ROC models that predict continuous curves extending to (1,1), imply zero search performance.

Lesion-classification performance is measured by the AUC value corresponding to the parameter. Lesion-classification performance is the ability to discriminate between LLs and NLs, not between diseased and non-diseased cases: the latter is measured by RSM-AUC. There is a close analogy between the two ways of measuring lesion-classification performance and CAD used to find lesions in screening mammography vs. CAD used in the diagnostic context to determine if a lesion found at screening is actually malignant. The former is termed CADe, for CAD detection, which in my opinion, is slightly misleading as at screening lesions are found not detected (“detection” is “discover or identify the presence or existence of something”, correct localization is not necessarily implied; the more precise term is “localize”). In the diagnostic context one has CADx, for CAD diagnostic, i.e., given a specific region of the image, is the region malignant?

Search and lesion-classification performance can be used as “diagnostic aids” to optimize performance of a reader. For example, is search performance is low, then training using mainly non-diseased cases is called for, so the resident learns the different variants of non-diseased tissues that can appear to be true lesions. If lesion-classification performance is low then training with diseased cases only is called for, so the resident learns the distinguishing features characterizing true lesions from non-diseased tissues that fake true lesions.

Finally, evidence for the RSM is summarized. Its correspondence to the empirical Kundel-Nodine model of visual search that is grounded in eye-tracking measurements. It reduces in the limit of large , which guarantees that every case will yield a decision variable sample, to the binormal model; the predicted pdfs in this limit are not strictly normal, but deviations from normality would require very large sample size to demonstrate. Examples were given where even with 1200 cases the binormal model provides statistically good fits, as judged by the chi-square goodness of fit statistic, Table 17.2. Since the binormal model has proven quite successful in describing a large body of data, it satisfying that the RSM can mimic it in the limit of large . The RSM explains most empirical results regarding binormal model fits: the common finding that b < 1; that b decreases with increasing lesion pSNR (large and / or ); and the finding that the difference in means divided by the difference in standard deviations is fairly constant for a fixed experimental situation, Table 17.3. The RSM explains data degeneracy, especially for radiologists with high expertise.

The contaminated binormal model2-4 (CBM), Chapter 20, which models the diseased distribution as having two peaks, one at zero and the other at a constrained value, also explains the empirical observation that b-parameter < 1 and data degeneracy. Because it allows the ROC curve to go continuously to (1,1), CBM does not completely account for search performance – it accounts for search when it comes to finding lesions, but not for avoiding finding non-lesions.

I do not want to leave the impression that RSM is the ultimate model. The current model does not predict satisfaction of search (SOS) effects27-29. Attempts to incorporate SOS effects in the RSM are in the early research stage. As stated earlier, the RSM is a first-order model: a lot of interesting science remains to be uncovered.

32.5.1 The Wagner review

The two RSM papers12,13 were honored by being included in a list of 25 papers the “Highlights of 2006” in Physics in Medicine and Biology. As stated by the publisher: "I am delighted to present a special collection of articles that highlight the very best research published in Physics in Medicine and Biology in 2006. Articles were selected for their presentation of outstanding new research, receipt of the highest praise from our international referees, and the highest number of downloads from the journal website.

One of the reviewers was the late Dr. Robert (“Bob”) Wagner – he had an open-minded approach to imaging science that is lacking these days, and a unique writing style. I reproduce one of his comments with minor edits, as it pertains to the most interesting and misunderstood prediction of the RSM, namely its constrained end-point property.

I’m thinking here about the straight-line piece of the ROC curve from the max to (1, 1). 1. This can be thought of as resulting from two overlapping uniform distributions (thus guessing) far to the left in decision space (rather than delta functions). Please think some more about this point–because it might make better contact with the classical literature. 2. BTW – it just occurs to me (based on the classical early ROC work of Swets & co.) – that there is a test that can resolve the issue that I struggled with in my earlier remarks. The experimenter can try to force the reader to provide further data that will fill in the space above the max point. If the results are a straight line, then the reader would just be guessing – as implied by the present model. If the results are concave downward, then further information has been extracted from the data. This could require a great amount of data to sort out–but it’s an interesting point (at least to me).

Dr. Wagner made two interesting points. With his passing, I have been deprived of the penetrating and incisive evaluation of his ongoing work, which I deeply miss. Here is my response (ca. 2006):

The need for delta functions at negative infinity can be seen from the following argument. Let us postulate two constrained width pdfs with the same shapes but different areas, centered at a common value far to the left in decision space, but not at negative infinity. These pdfs would also yield a straight-line portion to the ROC curve. However, they would be inconsistent with the search model assumption that some images yield no decision variable samples and therefore cannot be rated in bin ROC:2 or higher. Therefore, if the distributions are as postulated above then choice of a cutoff in the neighborhood of the overlap would result in some of these images being rated 2 or higher, contradicting the RSM assumption. The delta function pdfs at negative infinity are seen to be a consequence of the search model.

One could argue that when the observer sees nothing to report then he starts guessing and indeed this would enable the observer to move along the dashed portion of the curve. This argument implies that the observer knows when the threshold is at negative infinity, at which point the observer turns on the guessing mechanism (the observer who always guesses would move along the chance diagonal). In my judgment, this is unreasonable. The existence of two thresholds, one for moving along the non-guessing portion and one for switching to the guessing mode would require abandoning the concept of a single decision rule. To preserve this concept one needs the delta functions at negative infinity.

Regarding Dr. Wagner’s second point, it would require a great amount of data to sort out whether forcing the observer to guess would fill in the dashed portion of the curve, but I doubt it is worth the effort. Given the bad consequences of guessing (incorrect recalls) I believe that in the clinical situation, the radiologist will never guess. If the radiologist sees nothing to report, nothing will be reported. In addition, I believe that forcing the observer, to prove some research point, is not a good idea.

32.6 References

  1. Chakraborty DP. Computer analysis of mammography phantom images (CAMPI): An application to the measurement of microcalcification image quality of directly acquired digital images. Medical Physics. 1997;24(8):1269-1277.
  2. Chakraborty DP, Eckert MP. Quantitative versus subjective evaluation of mammography accreditation phantom images. Medical Physics. 1995;22(2):133-143.
  3. Chakraborty DP, Yoon H-J, Mello-Thoms C. Application of threshold-bias independent analysis to eye-tracking and FROC data. Academic Radiology. 2012;In press.
  4. Chakraborty DP. ROC Curves predicted by a model of visual search. Phys Med Biol. 2006;51:3463–3482.
  5. Chakraborty DP. A search model and figure of merit for observer data acquired according to the free-response paradigm. Phys Med Biol. 2006;51:3449–3462.

References

Dorfman, D. D., and E. Alf. 1969. “Maximum-Likelihood Estimation of Parameters of Signal-Detection Theory and Determination of Confidence Intervals - Rating-Method Data.” Journal Article. Journal of Mathematical Psychology 6: 487–96.

Dorfman, D. D., and K. S. Berbaum. 2000. “A Contaminated Binormal Model for ROC Data: Part Ii. A Formal Model.” Journal Article. Acad Radiol. 7 (6): 427–37. https://doi.org/10.1016/S1076-6332(00)80383-9.

Dorfman, D. D., K. S. Berbaum, C. E. Metz, R. V. Lenth, J. A. Hanley, and H. Abu Dagga. 1997. “Proper Receiving Operating Characteristic Analysis: The Bigamma Model.” Journal Article. Acad. Radiol. 4 (2): 138–49. https://doi.org/10.1016/S1076-6332(97)80013-X.

Metz, C. E., and X. Pan. 1999. “Proper Binormal ROC Curves: Theory and Maximum-Likelihood Estimation.” Journal Article. J Math Psychol 43 (1): 1–33.