Research 191 (2004) 49–58 www.elsevier.com/locate/heares

Tonotopic organization of the human probed with frequency-modulated tones

Nathan Weisz *, Christian Wienbruch, Sandra Hoffmeister, Thomas Elbert

Department of Psychology, University of Konstanz, Box D25, 78457 Konstanz, Germany

Received 31 July 2003; accepted 9 January 2004

Abstract

Using neuromagnetic source imaging, we investigated tonotopic representation and direction sensitivity in the auditory cortex of humans (N ¼ 15). For this purpose, source analysis was undertaken at every single sampling point during the presentation of a frequency-modulated tone (FM) sweeping slowly downward or upward across periods of 3 s duration. Stimuli were selected to target response properties of the central part of the primary auditory cortical field, which has been shown to exhibit sensitivity to distinct FM-sound features as compared to the ventral and dorsal part. Linear mixed-effects model statistics confirm tonotopic gradients in medial–lateral and anterior–posterior directions. The high resolution provided by this method revealed that the rela- tionship between frequency and spatial location of the responding neural tissue is nonlinear. The idea that neurons specifically sensitive to the employed sound characteristics (slow, downward modulation) were activated is supported by the fact that the upward sweep of identical duration produced a different pattern of functional organisation. Ó 2004 Elsevier B.V. All rights reserved.

Keywords: Tonotopy; MEG; High-resolution; LME

1. Introduction fields exist that process auditory information, reflecting the hierarchical and parallel processing architecture of A fundamental organizational principle of the central the auditory cortex. The exact number of fields is un- is the frequency specific (tonotopic) known for humans (Ehret, 1997; Talavage et al., 2000; representation of neuronal populations (de Ribaupierre, Wallace et al., 2002). However, one can expect the sit- 1997; Horikawa et al., 2001; Schreiner et al., 2000), uation to be similar to non-human primates, for which preserving the arrangement of hearing receptors in the Kaas et al. (1999) proposed a rough subdivision of the inner ear (high-to-low frequency representation from auditory cortex into three areas: the core (primary au- base to apex). This means that neurons processing in- ditory cortex; incorporating the primary auditory cor- formation from neighbouring hair cells are also neigh- tical field, AI), the belt (secondary auditory cortex; bours on a tonotopic map. On a cortical level, several including the secondary auditory cortical field, AII), and the parabelt (Kaas et al., 1999). In general, the degree of tonotopy decreases and the degree of multimodal inte- gration increases in the order mentioned above. * Corresponding author. Tel.: +49-7531-88-4606/4612; fax: +49- 7531-88-4061/4601. Tonotopy is most frequently investigated by mea- E-mail address: [email protected] (N. Weisz). suring neuronal responses to various stimuli consisting Abbreviations: AI , primary auditory cortical field; AII, secondary of a single carrier frequency, typically a pure tone. In auditory cortical field; AIC, Akaike information criterion; ANOVA, humans, magnetencephalography (MEG; e.g., Elbert analysis of variance; CF, characteristic frequency; ECD, equivalent et al., 2002; Pantev et al., 2003) has been a frequently current dipole; FM, frequency modulated; GOF, goodness of fit; LME, linear mixed effects modelling; LRT, likelihood ratio test; MEG, employed technique for noninvasively investigating the magnetencephalography; ML, maximum like; RCF, rate of change of functional organisation and reorganisation of the audi- frequency; SL, sensation level; SPL, sound pressure level tory cortex (for functional neuroimaging approaches see

0378-5955/$ - see front matter Ó 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.heares.2004.01.012 50 N. Weisz et al. / Hearing Research 191 (2004) 49–58 e.g., Lauter et al., 1985; Lockwood et al., 1999; Wes- and Cynader, 1985; Mendelson et al., 1993; Orduna singer et al., 2001). et al., 2001; see Tian and Rauschecker (1994) for an The conventional MEG-approach attempts to model investigation of neuronal responses to FM-sweeps in the the magnetic signals measured from the surface of the anterior auditory field). The dorsal and ventral stripe head by modelling the underlying source with a single show a directly opposite response pattern. Direction and equivalent current dipole (ECD) per hemisphere (see e.g. speed preference has been attributed to a common Elbert, 1998; Ham€ al€ ainen€ et al., 1993). A frequency- mechanism (Gordon and OÕNeill, 1998): an asymmetric dependent spatial shift in the location of the ECD is distribution of inhibitory sidebands of an excitatory considered as an indicator for tonotopic organization. tuning curve around a CF. Heil et al. (1992c) were able This strategy has been frequently employed for various to demonstrate an orderly arrangement of FM direction neuromagnetic components, e.g. the N1 (Godey et al., selective neurons in A1 that parallels the characteristic 2001; Lutkenh€ oner€ and Steinstrater, 1998; Pantev et al., frequency: regions sensitive to low tones preferred up- 1988, 1998), the Pa (Pantev et al., 1995) the steady state ward sweeps, whereas neurons with a high characteristic field (SSF) (Engelien et al., 2000; Pantev et al., 1996; Ross frequency seem to prefer downward sweeps. These re- et al., 2000) and the sustained field (Pantev et al., 1994). sults were recently corroborated by Zhang et al. (2003). However, this approach has important shortcomings: Of interest is also the observation that neurons in the 1. In order to gain an acceptable signal-to-noise ratio central stripe have the narrowest frequency tuning the same stimulus must be presented at a slow rate curves and the lowest scatter of CF corresponding to the (less than 1/s) and with many repetitions, thus impos- greatest frequency selectivity on a single- and multiunit ing constraints on the number of frequencies that can level (Ehret, 1997; Heil et al., 1992a; Schreiner et al., be investigated (usually only three or four). 2000). This means that sweep ranges exceeding the ex- 2. The power of the peak is in the frequency range citatory bandwidth of the frequency tuning curve at a below 12–15 Hz, i.e., the signal is contaminated by certain CF should lead to a corresponding spatial shift coherent biological noise that significantly distorts in the focal point of neuronal activity along the tono- source modelling. topic gradient. The observation of neuronal activation 3. Natural auditory stimuli to be processed by the audi- being triggered when the sweep crosses the CF of the tory cortex are generally complex and time-variant. neuronal cluster has been shown in animal experiments Since the study of complex tones is a difficult endeav- (Nelken and Versnel, 2000a,b). More specifically, re- our the use of amplitude and frequency modulated sponses seem to be evoked by modulations towards the tones poses an attractive alternative (Tian and Raus- CF of neurons, i.e., before instantaneous frequency checker, 1994): they are not as artificial as pure tones, reached CF, and not by modulations away from the CF yet their physical dimensions are more easily con- (Heil et al., 1992b). Employing a voltage sensitive dye, trolled than natural sounds. Horikawa et al. (1998) were able to show a spot-like The third point is reflected in the fact that FM tones response crossing the isofrequency bands in the primary provoke stronger neuronal responses than responses to auditory cortex, thus ÔtravelingÕ along the tonotopic the respective characteristic frequencies (CF) (deCharms axis. So far, a similar documentation of spatiotemporal et al., 1998; Nelken et al., 1994). deCharms et al. (1998) activation induced by FM-stimuli has not been at- demonstrated, for example, that only 5% of investigated tempted in humans. The present study exploited the neurons in AI are exclusively responsive to a single excellent temporal resolution capacity of MEG (Elbert, frequency region. The great majority of neurons show 1998; Ham€ al€ ainen€ et al., 1993) in order to investigate sensitivity to additional features. This is also reflected in whether ECD sources alter their location in a systematic the organization of the tonotopic map of AI in isofre- fashion during the processing of FM-sweeps. Data quency stripes (dorsal, central and ventral) at similar CF concerning neuronal coding of sweep rate and direction as shown in animal data (Ehret, 1997; Schreiner et al., stem essentially from work on nonhuman primates (e.g., 2000). Each stripe combines a different mixture of cats, bats, etc.). In the process of designing our stimuli, response characteristics to certain features of acous- efforts were undertaken that they matched response tic stimuli (binaural interaction, monotonicity of rate- preferences of the central stripe of AI as described by intensity function, dynamic range, etc.). Ehret (1997). Therefore, sweep direction was down- For frequency-modulated tones (FM-sweeps), i.e. wards from 8 to 0.5 kHz with an slow modulation fre- tones with a constantly changing instantaneous fre- quency of 1/6 Hz (downward slope of sinusoidal FM quency, these features are mainly direction, speed and tone; corresponding to an average frequency change of range. Concerning direction and speed the central iso- 2.5 kHz/s). This – for animal standards – unusually slow frequency stripe shows distinct response properties with speed of modulation spanning across 3 s was mainly a majority of neurons (about 2/3) preferring downward motivated by methodological reasons: (a) It cannot be sweeping tones (i.e., higher to lower frequencies) and expected that magnetic source imaging can adequately slow modulation speeds (Heil et al., 1992b; Mendelson resolve shifts in activation along a tonotopic gradient, N. Weisz et al. / Hearing Research 191 (2004) 49–58 51 when changes become widely dispersed due to a fast 2.2. Neuromagnetic recordings spread of activation. (b) A longer sweep duration allows a low-pass filter that eliminates spontaneous coherent Neuromagnetic data were recorded (sampling rate: noise (alpha activity) and leads to a greater number of 678.17 Hz; 0.1–100 Hz analogue filter) using a 148 data points that enter the statistics (see below). The es- channel whole-head magnetometer (4D Neuroimaging timation of tonotopic gradients will be more robust Inc., San Diego). Vertical and horizontal eye-move- against outlying values. A control condition kept ev- ments (EOG) were measured from above and below the erything constant except for sweep direction, which went eye and from the outer canthi. Epochs contaminated by upward from 0.5 to 8 kHz (upward slope of sinusoidal eye-movements or eye-blinks were not included in the FM tone). This condition was expected to yield con- averaged brain responses. siderably worse localizations than the experimental condition due to the fact that sound features were not 2.3. Procedure optimized for a presumed region equivalent to AI in animal studies. Auditory stimuli were monaurally presented via a Various criteria concerning localization and goodness plastic tube attached to an earpiece in the left ear of the of fit (GOF) were applied to the individual data (see participant. Individual sensation levels (SL) were as- method section for criteria) to exclude source localiza- sessed for four pure tones (500 ms; frequencies: 500, tions of low quality. From 15 original datasets only 2 1000, 3750 and 8000 Hz). The experimental stimuli were had to be excluded for the downward sweep condition. presented at intensities approximately 50 dB above SL. For the upward sweep condition, however, data gener- During the neuromagnetic recording session, subjects ally could not be modelled by a single source per laid in a supine position while watching a film of their hemisphere, indicating that source configurations for choice that was projected onto the ceiling of the mag- this case were complex (12 of the 15 subjects). Conse- netically shielded room. quently, statistical analysis was only performed for the The frequency-modulated (FM) tones were generated downward sweep data. with Matlab (version 5.2; The MathWorks, Inc.) ac- A linear mixed-effects model (LME) approach (Pin- cording to the formula by (Hartmann, 1998): heiro and Bates, 2000; see also Goldstein, 2003) was xðtÞ¼sinðx t þð sinðx t þ /ÞÞÞ; used to specify fixed- (i.e., average population charac- c m teristics) and random-effects (i.e., subject variability). where xðtÞ denotes the amplitude x at time t, xc is the This statistic is frequently employed in studies with re- constant carrier frequency, xm the modulation frequency, peated measures: among the advantages of this ap- / instantaneous phase and b the modulation index. proach are that it can (a) treat continuous data, (b) the Two blocks of FM-tones (3000 ms duration; 275 amount of data may vary among subjects (in our case: a epochs; 4000–4200 ms random SOA) were presented to different number of acceptable dipole fits per person) the participants in a counterbalanced order. The FM- and (c) it returns coefficients for the terms in our models tones in the blocks differed in their sweep-direction: In (fixed-effects). To test the coefficients of the fixed-effects the upward-condition they changed from 0.5 to 8 kHz parts, different measures were used (i.e., conditional t- (upward ramp of frequency), in the downward-condi- test for linear regression models). Furthermore, models tion from 8 to 0.5 kHz, respectively (downward ramp). (i.e., different fixed effects described in Section 2) were Due to the sinusoidal modulation, the sweep was compared regarding which one is more appropriate in changing the frequency at slower rates during the be- describing the data (Akaike information criterion [AIC] ginning and during the end of the stimulus (approx. 2 and likelihood ratio test [LRT]). For a detailed account kHz/s) but had a faster rate of change in the middle part of the LME approach, the reader is referred to Pinheiro (approx. 3.5 kHz/s). and Bates (2000). 2.4. Data-analysis

2. Methods Epochs of 4000 ms length (100 ms baseline) were ex- tracted from the raw data. Artefact free epochs, defined 2.1. Participants as EOG and MEG amplitudes below 100 lV or 5 pT, respectively, were averaged and filtered with a 0.7 Hz Fifteen subjects (age: range ¼ 17–32 years, mean ¼ 26, digital low-pass (Butterworth characteristic of 5th order) SD ¼ 4.21; 9 males) with normal hearing and no history in order to obtain the continuous response along the to- of an audiological or neurological disorder gave written notopic gradient of an auditory field(s) sensitive to the informed consent to participate in the study. They re- stimulus features applied. For each subject a group of ceived 15 for participation. The study was approved by 25–38 magnetic sensors was determined over the right the Konstanz University Ethical Review Board. hemisphere that covered best the evoked magnetic fields 52 N. Weisz et al. / Hearing Research 191 (2004) 49–58 from one hemisphere. Single ECDs were fitted for every calculation latency was centered at medium latency and sampling point during tone presentation (i.e., 0–3000 ms). scaled by factor 1000. Parameter estimation was per- Solutions were discarded from further analysis if their formed using maximum-likelihood (ML). Significance GOF was below 0.90, their localizations (BTI-coordi- of individual coefficients was tested using conditional t- nates) was more lateral than )7 cm or more medial than tests for linear regression models. Another procedure )1.5 cm, and more posterior and anterior than )4 or 4 cm, tested whether or not a better fit of the data might be a respectively. Furthermore, a dataset was only considered trivial effect of our model being too general (incorpo- as qualitatively sufficient when solutions meeting the de- rating too many terms). Models were therefore com- scribed criteria could be found for the majority of sam- pared among another with (a) the AIC, which penalizes pling points (i.e., >50%). This left 13 participants for the models with too many parameters (this leads to higher downward-sweep condition and 3 for the upward sweep. AIC values; i.e., models with smaller AIC are to be Because the downward-sweep FM was the only preferred), and (b) the LRT which considers the likeli- condition in which a sufficiently high number of results hood of a more general (L2) and a more restricted model as well as stable results were acquired, inferential sta- (L1). For 2 logðL2=L1Þ under the assumption that the tistics were only applied to this condition. For this restricted model is more adequate, the distribution of 2 purpose we chose an inside-out approach: In a first ex- the LRT-statistic is a v -distribution with k2–k1 degrees ploratory step, orthogonal polynomials of degree 1–5 of freedom (k being the number of parameters estimated were fitted to the individual data. This was done to de- in the respective models). LME analysis was done using cide which terms should enter the fixed-effects part of a the nlme-library (Pinheiro and Bates, 2000) of R (Ross LME model (Pinheiro and Bates, 2000) being used for and Gentleman, 1996) (version 1.6.2) for Mac OS X. the group statistics. Starting with a single linear term, the gain in explained variance was observed when add- 3. Results ing additional terms of the next higher polynomial. The five models tested were: The mean (SE) GOF for the five regression models is depicted in Fig. 1, which shows that a large amount of (1) y I ax, ¼ þ variance is explained by a single linear term (0.32–0.42; (2) y I ax bx2, ¼ þ þ according to CohenÕs (1988) classification a R2 larger (3) y I ax bx2 cx3, ¼ þ þ þ than 0.25 represents a substantial effect size). Repeated (4) y I ax bx2 cx3 dx4, ¼ þ þ þ þ measure ANOVAs indicate a significant effect for the (5) y ¼ I þ ax þ bx2 þ cx3 þ dx4 þ ex5, factor polynomial degrees (F4;48 for medial–lateral: where y is the response variable (i.e., the x-, y-andz- 14.53, for posterior–anterior: 19.31 and inferior–supe- direction), x the covariate (here: latency) and a–e the rior: 75.80; all p < 0:001). Post-hoc analysis showed coefficients. statistically significant and practically important in- Two criteria were introduced in order to select models creases in R2 in several cases: for subsequent statistics and to keep assumptions con- – Medial–lateral: The addition of a quadratic term to a cerning the parameters of the model to a minimum, i.e., to model with one linear term leads to a R2-gain of 0.23 determine a cut-off beyond it makes little sense to add (critical difference: 0.14). A further cubic term does additional terms: not contribute any significant extra information (R2- 1. A repeated measure ANOVA with the factor degrees gain: 0.04). of polynomials was calculated for the dependent vari- – Posterior–anterior: Same pattern as for medial–lateral able R2. In case of a significant effect, post-hoc analysis with a large R2-gain when adding a quadratic term (Tukey–Kramer procedure) was performed to show (0.23; critical difference: 0.17) and a statistically not which comparisons were statistically different, begin- significant for a further cubic term (0.13). ning with a first order polynomial moving upward in – Inferior–superior: Enormous R2-gain when adding a one-degree steps (i.e., comparison: degree 1 vs. 2, 2 quadratic term (0.40; critical difference: 0.11). A sta- vs. 3, etc.). Comparisons ceased when adding another tistically significant increase was also observed for term did not enhance R2 in a significant manner. an additional cubic term (0.13). 2. In addition, CohenÕs (1988) effect size criteria served Based on these results two fixed-effects models were for judging whether R2-enhancements were practi- specified for the LME statistic for the medial–lateral and cally important. For this, each additional term had posterior–anterior direction: to raise overall GOF by at least 0.09 (corresponding (1) Model 1: y ¼ I þ ax, to a medium effect size). (2) Model 2: y ¼ I þ ax þ bx2, This led to the selection of models 1–3. These were entered separately as fixed effects of a linear mixed where y is the response variable (x-, y-andz-axis), I the effects model statistic (LME). Random effects were de- intercept, x the covariate (latency) and a and b the fined as subject variability for the parameters. Prior to coefficients. N. Weisz et al. / Hearing Research 191 (2004) 49–58 53

The observed data for each individual (P1–P13) are plotted as red dots for each coordinate in Fig. 2. Due to the large amount of data only every 100th point is shown for the sake of clarity (the following LME statistic however include all data points). Additionally, data as predicted by the three models are depicted for each in- dividual (Model 1: blue dots; Model 2: green dots; Model 3: purple dots). Two striking features can be noted from visual inspection of the data: First of all, there is a very high inter-subject variability. Secondly, the application of a linear fixed-effects model (i.e., Model 1; blue dots in Fig. 2) does not seem appropriate in several cases. The latter impression is confirmed by the LME sta- tistic (Table 1). Both statistical criteria, the AIC and LRT, indicate the superiority of models in which non- linear (quadratic and cubic) terms are added (i.e., lower AIC- and higher LRT-values). Concerning the individ- ual fixed-effects terms in the medial–lateral case, the linear part remains statistically significant regardless of the model tested ()0.32 and )0.36; p < 0:05). The neg- ative sign of the coefficient means that overall source localization progressed laterally (towards the right) with time (model 2: from )4.10 to )5.17 cm). For the pos- terior–anterior direction, there was no significant linear term, but a highly significant quadratic one ()0.37; p < 0:01). The negative sign of the quadratic fixed-effect part indicates a spatial shift of sources further anterior up to the middle of the time-window (model 2: from )0.05 to 1.14 cm) and a shift back towards posterior localizations from the middle until the end model 2: 0.65 cm. For the inferior–superior axis only the linear fixed- effect of model 3 approaches significance. In this case, the negative sign implies a further inferior source localiza- tion as a function of time. More specifically, as the nonlinear terms must be taken into account, the repre- sentations first shift superiorally up to about 980 ms (model 3: from 5.03 to 5.27 cm) and then towards inferior (model 3: from 5.27 to 4.64 cm). To gain an impression of this ÔpopulationÕ-tonotopy, we plotted source locations resulting from the LME-statistics (i.e., the coordinates derived from the fixed effect) on a standard brain (Fig. 3).

4. Discussion

Concerning the tonotopic gradients, the models in this investigation are in accord with previous studies which Fig. 1. Goodness of fit as a function of the number of polynomial pa- have reported in most cases a medial to lateral and rameters. Plots show mean (standard error) goodness of fit for in- sometimes also a posterior to anterior shift of neural creasingly complex polynomial fits (see models 1–5 in Section 2) for the activity as the frequency decreases (Ehret, 1997; Elber- three spatial directions (medial–lateral, (a); posterior–anterior, (b); ling et al., 1982; Engelien et al., 2000; Kaas et al., 1999; inferior–superior, (c)). Pantev et al., 1989; Pantev et al., 1994; Pantev et al., 1996; For the inferior–superior direction a third model was Pantev et al., 1988; Pantev et al., 1995; Romani et al., tested next to the other two described above: 1982). Generally, such changes are observed for a subset of the sample only. The present study reveals the reasons: 2 3 (3) Model 3: y ¼ I þ ax þ bx þ cx . as the present analysis does not just ‘‘map’’ three or four 54 N. Weisz et al. / Hearing Research 191 (2004) 49–58

Fig. 2. Comparison of observed data and predicted values. Large box depicts the median value at each sampling point included in the statistic. Small boxes include the observed data (red dots) and the values predicted by the different models (model 1: blue dots; model 2: green dots; model 3: purple dots) each individual (P1–P13) by the LME statistic. Visualization was done for each direction separately ((medial–lateral, (a); posterior–anterior, (b); inferior–superior, (c)). For better visibility only every 100th point is shown. N. Weisz et al. / Hearing Research 191 (2004) 49–58 55

Table 1 Results of the LME statistic the three coordinate axes Term Coefficient SE t-value p-value AIC LRT Medial–lateral Model 1 I )4.733032 0.1589982 )29.767826 <0.0001 22250.785 Model 1 vs 2: x )0.315761 0.1567395 )2.014559 0.044 v2ð4Þ¼13031:5; Model 2 I )4.75709 0.2136012 )22.270894 <0.0001 9227.288 p < 0:0001 x )0.357949 0.1605976 )2.228857 0.0258 x2 0.054859 0.1469651 0.373279 0.7089 Posterior–anterior Model 1 I 0.9336182 0.150636 6.197843 <0.0001 26204.682 Model 1 vs 2: x 0.2564974 0.1696471 1.511947 0.1306 v2ð4Þ¼17089:22; Model 2 I 1.1360526 0.1613489 7.040967 <0.0001 9123.458 p < 0:0001 x 0.2350117 0.1775358 1.323742 0.1856 x2 )0.372458 0.1405524 )2.649959 0.0081 Inferior–superior Model 1 I 5.14172 0.2694313 19.083603 <0.0001 20538.728 Model 1 vs 2: x )0.209492 0.1271897 )1.647085 0.0996 v2ð4Þ¼15461:89; Model 2 I 5.217282 0.2591373 20.133273 <0.0001 5084.843 p < 0:0001 x )0.212659 0.1398734 )1.520368 0.1284 x2 )0.178403 0.1605133 )1.111455 0.2664 Model 3 I 5.211885 0.2532798 20.577574 <0.0001 )11544.417 Model 2 vs 3: x )0.198369 0.1020508 )1.943831 0.0519 v2ð5Þ¼16639:26; x2 )0.167005 0.1752601 )0.952899 0.3407 p < 0:0001 x3 0.030126 0.1777833 0.169453 0.8654 SE: standard error.

Fig. 3. MRI-Overlay of the fixed. 1822 dipoles are plotted in total. ÔColdÕ colors (blue) indicate positions at early latencies (beginning at 124 ms post stimulus onset; corresponding to high frequencies). ÔWarmÕ colors (pink) show positions at late latencies (ending 2809 ms post stimulus onset; corresponding to low frequencies). The cross indicates the location of the response to the stimulus onset (i.e., fit at 124 ms). Locations were obtained by using the coefficients of the most successful models (fixed effect) obtained by the LME statistic (see Table 1), i.e., this figure does not represent sources for a given individual but results from the group statistics. points but many for each subject the individual vari- roelectric and neuromagnetic studies of tonotopy. We ability becomes obvious as does a more complex, i.e., demonstrated for our results how the LME statistic ad- non-linear functional dependency between location and dresses this problem by taking random-effects into frequency. In addition, more inferior locations for low account. frequencies were indicated by the high resolution of the The question remains as to which parts and mecha- relationship. The considerable inter-individual variabil- nisms of the auditory cortex contributed to this finding. A ity has been noted in animal studies (Ehret, 1997) and trivial explanation, like the one suggesting that an overlap seems not surprising given the plasticity in representa- from N1 and sustained field contributed to the observed tional cortex (Pantev et al., 2003). Ehret (1997) therefore results, seems unlikely given that the time segment was argues that average statistics are inadequate and this in- three seconds. Moreover, comparable findings were not deed may pose a serious problem for several previous obtained for the upwards sweep. Generally, the specificity attempts to study tonotopy in humans. More recently, of the results for the downward sweep condition under- this question has also been recently raised by lines the importance of the stimulus features employed. Lutkenh€ oner€ et al. (2003) from the perspective of neu- It seems possible that the downward sweep has been 56 N. Weisz et al. / Hearing Research 191 (2004) 49–58 optimized for a certain area of AI, which exhibits this kind (3) It has to be kept in mind that the velocity of change of direction-preference with slow modulation rates (an- in this study was extremely slow, complicating compar- alogue to the central stripe in cat AI; Ehret, 1997). That isons with results from animal studies. In cases where the data of the upward sweep could not be modelled with such slow-going FM rates have been investigated (cat a single ECD per hemisphere is an indicator that at least studies), responsive neurons have been mainly found in two sources per hemisphere are involved, suggesting that the posterior auditory field (Heil and Irvine, 1998; Tian different source configurations were active for the and Rauschecker, 1998) and even then peak responses upward and for the downward sweep. This question were rather high compared to the rates used in the however cannot be solved with the currently available present investigation (<100 and 200 kHz/s, respectively). non-invasive methods. The limitations of the ECD anal- (4) Furthermore, it is difficult to ascribe unvarying ysis employed in the present study will be discussed below. response features to certain FM sweep characteristics Furthermore, direction-sensitivity and asymmetric (e.g. rate, range, direction) of auditory cortical neurons inhibitory side-bands of tuning curves seem to be related as they are also considerably dependent on other pa- (Gordon and OÕNeill, 1998; Nelken and Versnel, rameters. Heil and Irvine (1998) could show for neurons 2000a,b; Shamma et al., 1993). Opposite to FM-up- in the posterior auditory field of cats that e.g., the rate of wards-sensitive neurons, downwards-sensitive ones have change of frequency (RCF) varied with sound pressure stronger inhibitory side-bands on the low-frequency side level (SPL) of the stimulus. Furthermore, the directional of the tuning-curve. This means that a downward ori- sensitivity changed with RCF and SPL. Also, the in- ented sweep leads to a ÔproperÕ sequence of excitation stantaneous frequency leading to a response of the and inhibition when probing downward-sensitive neu- neuron depended on SPL and FM direction. leading to a rons. Upward oriented sweeps would first lead to an response of the neuron depending on SPL and FM di- inhibition of downward-sensitive neurons. These argu- rection. This outcome suggests that our results are de- ments (sweep direction/rate preference, asymmetry of pendent on the combination of the stimulus features and inhibitory side-bands and their relation to direction that further studies are needed before more general in- sensitivity) make a case that the downward-oriented terpretations become possible. sweep lead to a focal activation of the central part of AI The importance of our work lies in the fact that it is that shifted along the tonotopic axis. Other explanations the first study in humans to probe tonotopic represen- might arise, however, as we begin to understand more tation in a high-density mode. Using this mode we are about the cortical processing of FM-sweeps, which until able to make assertions that could not be made using a recently has been confined mainly to AI. conventional approach with only 3–5 carrier frequencies. There are several factors that pose limitations to the In addition, we expect that future parametric studies in interpretation of our results: humans will significantly expand our understanding of (1) While designing the study, we relied on animal, functional organisation of auditory representational mostly non-primate (cat) studies as reviewed by Ehret cortex by varying stimulus features of FM-tones, as well (1997). There are, however, strong species differences as by continuing to explore and apply statistical proce- (for review see Eggermont, 2001), so in the case of hu- dures which can accommodate nonlinear aspects and the mans it cannot be excluded that subdivisions of AI and considerable inter-subject variability (e.g. LME). their response properties might not exactly match the ones assumed in the introduction. Nevertheless, the finding of acceptable ECD fits for the downward sweep Acknowledgements for a wide time range point to a relatively focal activa- tion within the auditory cortex. Future, probably inva- This work was supported by the Deutsche Fors- sive studies, will have to resolve the question whether chungsgemeinschaft (DFG; EL 101/20). We thank Re- there exist isofrequency stripes within AI with a strong naud Lancelot and Willi Nagl for advice concerning preference for downward sweeps. linear mixed effects modelling, Christina Robert for (2) It has to be assumed that the processing of an helpful comments on an earlier version of the manu- FM tone will lead to activation in several auditory script, and Thomas Hartmann for assistance during fields, even though it is unlikely that the stimuli em- data acquisition. ployed lead to an equal and thus global activation in all fields (Eggermont, 2001). Thus source configura- tions can be too complex to be modelled by ECDs References and may be to transient in time to be tracked by hemodynamic imaging methods. Only, if a certain Cohen, J., 1988. Statistical Power Analysis for the Behavioral Sciences, second ed. Lawrence Erlbaum Associates, Hillsdale, NJ. cortical region dominates the pattern of neuromag- de Ribaupierre, F., 1997. Acoustical information processing in the netic activation, a tontotopic gradient appears. This auditory thalamus and . In: R, R., (Ed.), The central seems to be the case for the downward sweep. auditory system. Oxford University Press, New York, pp. 317–383. N. Weisz et al. / Hearing Research 191 (2004) 49–58 57 deCharms, R.C., Blake, D.T., Merzenich, M.M., 1998. Optimizing the normal human auditory system: responses to 0.5 and 4.0 kHz sound features for cortical neurons. Science 280, 1439–1443. tones at varied intensities. Cereb. Cortex 9, 65–76. Eggermont, J.J., 2001. Between sound and perception: reviewing the Lutkenh€ oner,€ B., Steinstrater, O., 1998. High-precision neuromagnetic search for a neural code. Hear. Res. 157, 1–42. study of the functional organization of the human auditory cortex. Ehret, G., 1997. The auditory cortex. J. Comp. Physiol. [A] 181, Audiol. Neurootol. 3, 191–213. 547–557. Lutkenh€ oner,€ B., Krumbholz, K., Seither-Preisler, A., 2003. Studies of Elberling, C.B.C., Kofoed, B., Lebech, J., Saermark, K., 1982. tonotopy based on wave N100 of the auditory evoked field are Auditory magnetic fields: source location and Ôtonotopical organi- problematic. Neuroimage 19, 935–949. zationÕ in the right hemisphere of the human brain. Scand. Audiol. Mendelson, J.R., Cynader, M.S., 1985. Sensitivity of cat primary 11, 61–65. auditory cortex (AI) neurons to the direction and rate of frequency Elbert, T., 1998. Neuromagnetism. In: Andra,€ W., Nowak, H. (Eds.), modulation. Brain Res. 327, 331–335. Magnetism in Medicine: A Handbook. Wiley-VCH, Berlin, pp. Mendelson, J.R., Schreiner, C.E., Sutter, M.L., Grasse, K.L., 190–261. 1993. Functional topography of cat primary auditory cortex: Elbert, T., Sterr, A., Rockstroh, B., Pantev, C., Muller, M.M., Taub, responses to frequency-modulated sweeps. Exp. Brain Res. 94, 65– E., 2002. Expansion of the tonotopic area in the auditory cortex of 87. the blind. J. Neurosci. 22, 9941–9944. Nelken, I., Versnel, H., 2000a. Responses to linear and logarithmic Engelien, A., Schulz, M., Ross, B., Arolt, V., Pantev, C., 2000. A frequency-modulated sweeps in ferret primary auditory cortex. combined functional in vivo measure for primary and secondary Eur. J. Neurosci. 12, 549–562. auditory cortices. Hear. Res. 148, 153–160. Nelken, I., Versnel, H., 2000b. Responses to linear and logarithmic Godey, B., Schwartz, D., de Graaf, J.B., Chauvel, P., Liegeois- frequency-modulated sweeps in ferret primary auditory cortex. Chauvel, C., 2001. Neuromagnetic source localization of Eur. J. Neurosci. 12, 549–562. auditory evoked fields and intracerebral evoked potentials: a Nelken, I., Prut, Y., Vaadia, E., Abeles, M., 1994. In search of the best comparison of data in the same patients. Clin. Neurophysiol. 112, stimulus: an optimization procedure for finding efficient stimuli in 1850–1859. the cat auditory cortex. Hear. Res. 72, 237–253. Goldstein, H., 2003. Multilevel Statistical Models, second ed. Edward Orduna, I., Mercado 3rd, E., Gluck, M.A., Merzenich, M.M., 2001. Arnold, London. Spectrotemporal sensitivities in rat auditory cortical neurons. Hear. Gordon, M., OÕNeill, W.E., 1998. Temporal processing across Res. 160, 47–57. frequency channels by FM selective auditory neurons can account Pantev, C., Hoke, M., Lehnertz, K., Lutkenhoner, B., Anogianakis, for FM rate selectivity. Hear. Res. 122, 97–108. G., Wittkowski, W., 1988. Tonotopic organization of the human Ham€ al€ ainen,€ M., Hari, R., Ilmoniemi, R.J., Knuutila, J., Lounasmaa, auditory cortex revealed by transient auditory evoked O.V., 1993. Magnetencephalography – theory, instrumentation, magnetic fields. Electroencephalogr. Clin. Neurophysiol. 69, 160– and applications to noninvasive studies of the working human 170. brain. Rev. Modern Phys. 65, 413–497. Pantev, C., Hoke, M., Lutkenhoner, B., Lehnertz, K., 1989. Tonotopic Hartmann, W.M., 1998. Signals, Sound, and Sensation. Springer, New organization of the auditory cortex: pitch versus frequency York. representation. Science 246, 486–488. Heil, P., Irvine, D.R., 1998. Functional specialization in auditory Pantev, C., Eulitz, C., Elbert, T., Hoke, M., 1994. The auditory evoked cortex: responses to frequency-modulated stimuli in the catÕs sustained field: origin and frequency dependence. Electroencepha- posterior auditory field. J. Neurophysiol. 79, 3041–3059. logr. Clin. Neurophysiol. 90, 82–90. Heil, P., Rajan, R., Irvine, D.R., 1992a. Sensitivity of neurons in cat Pantev, C., Bertrand, O., Eulitz, C., Verkindt, C., Hampson, S., primary auditory cortex to tones and frequency-modulated stimuli. Schuierer, G., Elbert, T., 1995. Specific tonotopic organizations of II: Organization of response properties along the ÔisofrequencyÕ different areas of the human auditory cortex revealed by simulta- dimension. Hear. Res. 63, 135–156. neous magnetic and electric recordings. Electroencephalogr. Clin. Heil, P., Rajan, R., Irvine, D.R., 1992b. Sensitivity of neurons in cat Neurophysiol. 94, 26–40. primary auditory cortex to tones and frequency-modulated stimuli. Pantev, C., Roberts, L.E., Elbert, T., Ross, B., Wienbruch, C., 1996. I: Effects of variation of stimulus parameters. Hear. Res. 63, Tonotopic organization of the sources of human auditory steady- 108–134. state responses. Hear. Res. 101, 62–74. Heil, P., Langner, G., Scheich, H., 1992c. Processing of frequency- Pantev, C., Weisz, N., Schulte, M., Elbert, T., 2003. Plasticity of the modulated stimuli in the chick auditory cortex analogue: evidence human auditory cortex. In: Pessoa, L., de Weerd, P. (Eds.), Filling- for topographic representations and possible mechanisms of In: From Perceptual Completion to Cortical Reorganization. rate and directional sensitivity. J. Comp. Physiol. [A] 171, 583– Oxford Univ. Press, Oxford. 600. Pinheiro, J.C., Bates, D.M., 2000. Mixed-Effects Models in S and S- Horikawa, J., Nasu, M., Taniguchi, I., 1998. Optical recording of Plus. Springer, New York. responses to frequency-modulated sounds in the auditory cortex. Romani, G.L., Williamson, S.J., Kaufman, L., Brenner, D., 1982. Neuroreport 9, 799–802. Characterization of the human auditory cortex by the neuromag- Horikawa, J., Hess, A., Nasu, M., Hosokawa, Y., Scheich, H., netic method. Exp. Brain Res. 47, 381–393. Taniguchi, I., 2001. Optical imaging of neural activity in multiple Ross, I., Gentleman, R., 1996. R: a language for data analysis and auditory cortical fields of guinea pigs. Neuroreport 12, 3335– graphics. J. Computat. Graphical Statist. 5, 299–314. 3339. Ross, B., Borgmann, C., Draganova, R., Roberts, L.E., Pantev, C., Kaas, J.H., Hackett, T.A., Tramo, M.J., 1999. Auditory processing 2000. A high-precision magnetoencephalographic study of human in primate cerebral cortex. Curr. Opin. Neurobiol. 9, 164– auditory steady-state responses to amplitude-modulated tones. J. 170. Acoust. Soc. Am. 108, 679–691. Lauter, J.L., Herscovitch, P., Formby, C., Raichle, M.E., 1985. Schreiner, C.E., Read, H.L., Sutter, M.L., 2000. Modular organization Tonotopic organization in human auditory cortex revealed of frequency integration in primary auditory cortex. Ann. Rev. by positron emission tomography. Hear. Res. 20, 199– Neurosci. 23, 501–529. 205. Shamma, S.A., Fleshman, J.W., Wiser, P.R., Versnel, H., 1993. Lockwood, A.H., Salvi, R.J., Coad, M.L., Arnold, S.A., Wack, D.S., Organization of response areas in ferret primary auditory cortex. J. Murphy, B.W., Burkard, R.F., 1999. The functional anatomy of Neurophysiol. 69, 367–383. 58 N. Weisz et al. / Hearing Research 191 (2004) 49–58

Talavage, T.M., Ledden, P.J., Benson, R.R., Rosen, B.R., Melcher, Wallace, M.N., Johnston, P.W., Palmer, A.R., 2002. Histochemical J.R., 2000. Frequency-dependent responses exhibited by identification of cortical areas in the auditory region of the human multiple regions in human auditory cortex. Hear. Res. 150, 225– brain. Exp. Brain Res. 143, 499–508. 244. Wessinger, C.M., VanMeter, J., Tian, B., Van Lare, J., Pekar, J., Tian, B., Rauschecker, J.P., 1994. Processing of frequency-modulated Rauschecker, J.P., 2001. Hierarchical organization of the human sounds in the catÕs anterior auditory field. J. Neurophysiol. 71, auditory cortex revealed by functional magnetic resonance imaging. 1959–1975. J. Cogn. Neurosci. 13, 1–7. Tian, B., Rauschecker, J.P., 1998. Processing of frequency-modulated Zhang, L.I., Tan, A.Y., Schreiner, C.E., Merzenich, M.M., 2003. sounds in the catÕs posterior auditory field. J. Neurophysiol. 79, Topography and synaptic shaping of direction selectivity in 2629–2642. primary auditory cortex. Nature 424, 201–205.