THE SEARCH BEHAVIOR OF SEA LAMPREY DURING THEIR NON-HOMING REPRODUCTIVE MIGRATION IN THE GREAT LAKES By Trevor D. Meckley A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Fisheries and Wildlife – Doctor of Philosophy 2015 ABSTRACT THE SEARCH BEHAVIOR OF SEA LAMPREY DURING THEIR NON-HOMING REPRODUCTIVE MIGRATION IN THE GREAT LAKES By Trevor D. Meckley Non-homing organisms are thought to rely on stable geophysical features (e.g., landslope), and encounter with distinct attributes of the environment (e.g., landmarks) in a predicable sequence to locate reproductive habitat (e.g., rivers always occur on a coastline). However little is known about how non-homing fishes complete a large-scale migration, as most research has focused on homing fishes that rely on geomagnetic cues to return to a natal spawning site. The invasive sea lamprey offers insight into non-homing migration, as individuals complete a single non-homing migration to rivers following translocation by host fishes in the Great Lakes. In this dissertation, Chapters 2 and 3, are devoted to developing a framework for describing animal behavior from telemetered observations and development of a standardized approach for assessing and filtering VEMCO Positioning System (VPS) data based on an estimate of horizontal position precision (HPE). This methodology was imperative for exploration of the sea lamprey migration with the underwater VPS telemetry technology. In Chapter 3, we described how sea lamprey orient to a coast when in a lake and hypothesized that sea lamprey navigate to the nearest coast by (1) orienting to the local bathymetric gradient and (2) maintain straight movements counter to the local slope to move towards shallow water. Three-dimensional (3-D) paths of migrating female sea lamprey were obtained by an acoustic array with 3 km2 of coverage, centered 3.3 km from the coast in Lake Huron. The findings of this chapter indicate that sea lamprey sampled an area of lake-bottom to assess absolute hydrostatic pressure and to select a heading towards reducing pressure (shallower water). In contrast to natal homing migrations, the sea lamprey appears consistent with nonhoming orientation to a general region with a simple set of rules based on local topography. Chapter 4 focused on the sea lamprey migration along a coastline and near a river mouth. Upon reaching a coastline, sea lampreys move parallel to shore. Prior studies indicated the presence of larval odor in river water increased the likelihood that a migrant entered a river. However, it was not known whether larval odor played a role in navigation (guiding the migrant to the river mouth) or mediated habitat selection by labeling the suitability of a river for spawning. In a two-year study using a 2 km2 acoustic array, the 3-D paths of sea lamprey were documented as they approached, entered, or bypassed the Ocqueoc River in northern Michigan, under one of two conditions: (1) low larval odor; and, (2) higher larval odor, created by increased larval abundance plus the addition of the synthetic larval odor components, petromyzonamine disulfate (PADS), petromyzosterol disulfate (PSDS), and petromyzonol sulfate (PZS) to a 1 x 10 -12 M concentration. A coupled hydrodynamic and dye concentration model predicted the hydraulic conditions experienced by each sea lamprey by estimating water conditions (velocity, temperature, etc.) at each fish position and allowed for assignment of whether a position was inside or outside of the Ocqueoc River plume. Encounter with river water appears to trigger localized search, regardless of larval odor content. However, when larval odor was abundant, the migrant was more likely to enter the river. Whether a migrant enters a river, is modulated by the presence of detectable larval odor, manipulation of river selection by invasive sea lamprey for management is viable in rivers with high encounter rates. Finally, Chapter 5 covers the implications for how altering migration routes in the sea lamprey via the application of synthesized pheromones contained in larval odor could be profitable for management. To my wife for five great years in Michigan… iv AKNOWLEDGEMENTS I would like to thank my advisor, Dr. C. Michael Wagner immensely for his contributions to my dissertation. His mentoring on research design, interpretation of results, and presentation of findings has made me a much improved scientist. The members of my graduate committee all supported different aspects of my dissertation and included Dr. James Miller, Dr. Daniel Hayes, Dr. Charles Krueger, and Dr. Weiming Li. I thank my committee, Dr. Eliezer Gurarie, Dr. Tom Binder, Dr. Christopher Holbrook, Dr. Frank Smith, and Jessica Barber for reviewing chapters. I would like to specifically thank Dr. Eliezer Gurarie and Dr. Christopher Holbrook for their collaboration and help on my dissertation, including teaching me the ways of R. Dr. Tuan Nguyen and Dr. Mantha S. Phanikumar are responsible for the hydrodynamic models provided in Chapter 4. I thank all personnel at the United States Geological Survey Hammond Bay Biological Station, Millersburg, Michigan, and the United States Fish and Wildlife Service Marquette Biological Station, Marquette, Michigan, for their assistance in animal capture, housing, maintenance, and technical support during these projects. In addition, I would like to thank specific individuals at the stations who have worked directly with me on my projects in some capacity including: Dr. Michael Hansen, Dr. Roger Bergstedt, Erick Larson, Karen Slaght, Jessica Barber, and Dr. Nicholas Johnson. I also received additional support from Glen Black, Tyler Buchinger, Cory Olaf Brant, Dr. Michael Siefkes, Dr. Michael Twohey, Dr. Darryl Horndorp, Captain Joseph Bergan, Dr. Robert Goodwin, Dr. John Hume, and Dr. Thomas Luhring. I would like to give special thanks to field technicians for their hard work, dedication, and patience during these projects: Eric Willman, Brett Diffin, Amber Masters, Greg Byford, Skye Fissette, Carrie Kozel, Sarah Ptasznik, and Jeff Yaklin. v Lastly, I thank my entire family. My mom, dad, Janet and Joel have provided incredible support for me during my Ph.D. degree. Most importantly I thank my wife, for all of her love and support and for moving to Michigan, even though I went away every year to chase “ugly” fish. vi TABLE OF CONTENTS LIST OF TABLES ........................................................................................................................ x LIST OF FIGURES ................................................................................................................... xiii CHAPTER 1 .................................................................................................................................. 1 AN ADAPTIVE WORKFLOW FOR STUDYING BEHAVIOR WITH TELEMETERED OBSERVATIONS. .................................................................................................................. 1 ABSTRACT ............................................................................................................................... 1 INTRODUCTION ..................................................................................................................... 1 GENERAL WORKFLOW ....................................................................................................... 4 Question Development ............................................................................................................. 4 Data Collection........................................................................................................................ 5 Data processing ....................................................................................................................... 6 Analyses ................................................................................................................................... 7 APPROACH AND IMPLEMENTATION ............................................................................. 9 Step 1: Question Development .............................................................................................. 11 Step 2: Data Collection ......................................................................................................... 12 Step 3: Data Processing ........................................................................................................ 13 Step 4: Analyses ..................................................................................................................... 14 CONCLUSION ........................................................................................................................ 14 APPENDIX .............................................................................................................................. 16 REFERENCES ........................................................................................................................ 20 CHAPTER 2 ................................................................................................................................ 25 A PROCESS FOR FILTERING HYPERBOLICALLY POSITIONED UNDERWATER TELMETRY DATA WITH HORIZONATAL POSITIONG ERROR (HPE) ..................... 25 ABSTRACT ............................................................................................................................. 25 INTRODUCTION ................................................................................................................... 26 Background ............................................................................................................................ 26 The VPS system ...................................................................................................................... 27 Filtering spatial data with HPE ............................................................................................ 28 METHODS .............................................................................................................................. 30 Error evaluation based on test data ...................................................................................... 31 HPE evaluation with test data ............................................................................................... 32 Filtering fish data .................................................................................................................. 34 RESULTS ................................................................................................................................. 34 Step 1: Establishing data quality objectives .......................................................................... 34 Step 2: Estimating the baseline position confidence in the array ......................................... 35 vii Step 3a: Evaluating HPE filters consistent with the data quality objectives and guides ...... 36 Criterion 1 ............................................................................................................................. 37 Criterion 2 ............................................................................................................................. 37 Criterion 3 ............................................................................................................................. 37 Criterion 4 ............................................................................................................................. 38 Step 3b: Selecting an HPE cutoff .......................................................................................... 38 Step 3c: Evaluating the potential for introduced bias through application of HPE cutoff ... 39 DISCUSSION .......................................................................................................................... 39 ACKNOWLEDGEMENTS .................................................................................................... 45 APPENDIX .............................................................................................................................. 46 REFERENCES ........................................................................................................................ 60 CHAPTER 3 ................................................................................................................................ 64 HOW DO NON-HOMING FISHES FIND THE SHORE? EVIDENCE FOR BATHYMETRIC ORIENTATION IN MIGRATING SEA LAMPREY ............................ 64 ABSTRACT ............................................................................................................................. 64 INTRODUCTION ................................................................................................................... 65 METHODS .............................................................................................................................. 68 RESULTS ................................................................................................................................. 80 DISCUSSION .......................................................................................................................... 83 ACKNOWLEDGEMENTS .................................................................................................... 90 APPENDIX .............................................................................................................................. 91 REFERENCES ...................................................................................................................... 149 CHAPTER 4 .............................................................................................................................. 158 DOES LARVAL SEA LAMPREY ODOR AID NAVIGATION OR GUIDE HABITAT SELECTION DECISIONS BY ADULTS .......................................................................... 158 ABSTRACT ........................................................................................................................... 158 INTRODUCTION ................................................................................................................. 159 METHODS ............................................................................................................................ 162 General Methods ................................................................................................................. 162 Study Site ............................................................................................................................. 163 Experimental Subjects ......................................................................................................... 163 Pheromone application........................................................................................................ 164 VPS Array Performance and data treatment ....................................................................... 165 Hydrological Data ............................................................................................................... 166 Identifying active sea lamprey ............................................................................................. 169 Characterizing river plume encounter and entry ................................................................ 170 Movement Patterns .............................................................................................................. 172 Effect of Synthetic Odor ....................................................................................................... 174 RESULTS ............................................................................................................................... 174 Sea Lamprey Activity ........................................................................................................... 174 Characterizing river plume encounter and river entry ....................................................... 175 viii Movement Patterns .............................................................................................................. 177 Effect of synthetic odor ........................................................................................................ 179 DISCUSSION ........................................................................................................................ 179 Movement outside of the river plume .................................................................................. 181 River plume encounter ......................................................................................................... 182 Movement in the river plume ............................................................................................... 184 The role of larval odor......................................................................................................... 185 River entry ........................................................................................................................... 187 Odor mediated manipulation of sea lamprey habitat selection ........................................... 188 Summary .............................................................................................................................. 189 ACKNOWLEDGEMENTS .................................................................................................. 190 APPENDIX ............................................................................................................................ 191 REFERENCES ...................................................................................................................... 291 CHAPTER 5 ............................................................................................................................. 298 ISOLATING OPPORTUNITIES FOR MANIPULATING SPAWNING HABITAT SELECTION OF INVASIVE SEA LAMPREY (PETROMYZON MARINUS) IN THE LAURENTIAN GREAT LAKES ....................................................................................... 298 ABSTRACT ........................................................................................................................... 298 INTRODUCTION ................................................................................................................. 298 RIVER ENTRY FRAMEWORK ........................................................................................ 300 River Plume Encounter ........................................................................................................ 300 River Entrance, Upstream Movement, and Tributary Selection ......................................... 301 River Retention .................................................................................................................... 304 MANIPULATING THE MIGRATION .............................................................................. 304 River Plume Encounter ........................................................................................................ 305 Cohort-Size Maintenance .................................................................................................... 305 Recruitment to habitat deficient for larval survival ............................................................ 306 Chemical barrier and Push-Pull ......................................................................................... 307 Larval Odor Removal (TFM) .............................................................................................. 308 CONCLUSION ...................................................................................................................... 309 APPENDIX ............................................................................................................................ 310 REFERENCES ...................................................................................................................... 317 ix LIST OF TABLES Table 1.1: Studies of animal behavior that focus on telemetry data should follow this general framework and report on each component: question development data collection, data processing, and analysis. Using the framework ensures that at the start of the study there is clear consideration of what data is necessary to properly answer the question and describe the behavior in an exploratory, explanatory or predictive context. ...................................................................17 Table 1.2: The components of Ethology. ......................................................................................18 Table 2.1: Filtering Objectives. The four specific criteria adopted to establish the data quality objectives for the project ................................................................................................................47 Table 2.2: Criteria for selection of an HPE filter cutoff ..............................................................48 Table 3.1: Sea lamprey were primarily detected on the nearest coast within 72 hours of release. This included 78 % of individuals that were observed stopping in the array and moving on the first night (n=22) and 61% of all individuals released (n=67). Note that Time to reach the receiver mean and min/max are only listed for the 22 individuals. Only 4 of 67 sea lamprey’s first detections occurred outside of the nocturnal movement period, with 3 of 4 occurring on the first day of release with 2 at Forty Mile point (East), 1 at the Ocquoec river mouth (South) and 1 near Cheobygran river mouth (West). ..........................................................................................92 Table 3.2: Individual t-tests for each sea lamprey of whether sea lamprey exhibited a persistence in turning (theta; -pi to pi), different from 0, revealed that more sea lamprey had a bias of left (negative) or right (positive) turns during phase 1 (P1), but not during phase 2 (P2). Most persistent turns were counterclockwise. The magnitude is also greater for those that are significant in phase 1 than phase 2. A pairwise t-test of the concentration in turning (Rho) for each individual between phase 1 and phase 2 revealed that individuals tended to go straighter during phase 2. Bold values are significant. .................................................................................93 Table 3.3: Pairwise t-tests to determine if there was a difference in the mean depth, standard deviation in depth, mean depth when maintaining vertical depth (Hz Mean Depth), or ground speed during phase 1(P1) and before the transition to phase 2 (P2). The tests support our observations that these two phases are different. Sea lamprey had greater mean depths and less variation during phase 1, and when moving at a particular depth in the water column, sea lamprey spent time on the bottom during phase 1 and closer to the surface during phase 2. The ground speed (GS) was also significantly lower during phase 1 than phase 2. The only individuals with less variation in depth during phase two were those individuals swimming mostly at the surface with occasional vertical excursions. Individuals T42-T56 did not have pressure sensitive tags. ..................................................................................................................94 x Table 4.1: Three specific criteria were adopted to establish the data quality objectives based on an extensive review of data quality (Meckley et al. 2014a). ......................................................192 Table 4.2: Criteria for selection of an HPE filter cutoff. In the data values listed for each criterion the HPE cutoffs are listed acceptable HPE cutoffs are listed for 2010 above and 2011 below. ..........................................................................................................................................193 Table 4.3: Akaike’s information criterion (AIC) values for linear mixed effects models using river entry (Em) as a binary response variable to various combinations of fixed effects including year (Y), iterative number of nights the sea lamprey encountered river water (N), a year by night number interaction (YxN), average nightly discharge of the Ocqueoc River (DC: m3·S-1), average lake temperature encountered (LT: C ̊), average nightly river temperature (RT; C ̊ ), the difference in experienced lake temperature and river temperature (DFT; C ̊ ), minimum distanced reached from the river mouth (D: km), and average water velocity experienced in the lake (V). ... ......................................................................................................................................................194 Table 4.4: Akaike’s information criterion (AIC) values for a nonlinear mixed-effects model in the formulation described by Lindstrom and Bates (1990), using the function “nmle” in R (R Development Core Team 2015). Ground speed (Vm) as a response variable to various combinations of fixed effects including year (Y), the log of the cumulative distance traveled (D), whether encounter with river water occurred yet in the path (E), or if the point was currently in the plume (P), temperature at fish position (T), and an interaction effect between water current magnitude and direction by fish heading interaction (CxF); and finally the random effect of individuals (animal ID). ..............................................................................................................195 Table 4.5: Akaike’s information criterion (AIC) values for a nonlinear mixed-effects model in the formulation described by Lindstrom and Bates (1990), using the function “nmle” in R (R Development Core Team 2015). Path straightness (Sm) as a response variable to various combinations of fixed effects including year (Y), the log of the cumulative distance traveled (D), if whether encounter with river water occurred yet in the path (E), or the point was currently in the plume (P); and finally the random effect of individuals (animal ID). Model 6 and 9 had the lowest AIC scores. ......................................................................................................................196 Table 4.6: To evaluate the effect of a host of variables on river entry (Em), a mixed effects logistic regression with a binary response variable (river entry: 1 or 0) was implemented in R using glmer (lme4 package, Bates et al. 2014). The best fitting models, as determined by AIC, for river entry included year and minimum distance reached from the river and either the lake temperature experienced (Model 10) or the difference in lake temperature experienced and the river temperature (Model 11). The fixed effects included year (Y), iterative number of nights encountering river water (N), a year by night number interaction(YxN), average nightly discharge of the Ocqueoc River (DC: m3·S-1), average lake temperature encountered (LT: C ̊), average nightly river temperature (RT; C ̊ ), the difference in experienced lake temperature and river temperature(DFT; C ̊ ), minimum distanced reached from the river mouth (D: km), and average water velocity experienced in the lake (V). Here is Model 11 output. ..........................197 xi Table 4.7: Output from the summary of the best fit nonlinear mixed-effects model for the continuous response variable ground speed (body lengths per second), Model 9. The model included the temperature experienced (T), whether encounter with the river plume had occurred in the path (E), whether the sea lamprey was in the river plume (P), and an interaction between water current and if the fish was moving towards or away from the current direction. The correlation statement accounts for first order autocorrelation in the residuals that occurred from analyzing multiple simultaneous steps relating to the same individual. The random effect of animal ID is included. .................................................................................................................199 Table 4.8: Output from the summary of the best fit nonlinear mixed-effects model for path straightness (0-1), Model 9. The model included the effect of the interaction between cumulative distance traveled (D) and Encounter with the river plume and the fixed effect of being in the river plume. The correlation statement accounts for first order autocorrelation in the residuals that occurred from analyzing multiple simultaneous steps relating to the same individual. The random effect of animal ID is included. The distance is divided by 1000 to put the variable in km rather than m for easier interpretation. ........................................................................................201 Table 5.1: Six general strategies for manipulating sea lamprey migration behavior are presented including the stimulus affected, how the stimulus is affected (action), the goal of the manipulation and the type of manipulation. Only option one does not stand alone as a management strategy as it manipulates behavior but not a decision to select habitat. ...............311 xii LIST OF FIGURES Figure 1.1: Analysis objectives aim to identify what is the animal doing (exploratory), why the animal is doing what it is doing (explanatory), and can we anticipate how it will change what it will do under different conditions (predictive)? The amount known about the animal often plays a role in where researchers aim study objectives initially and we identify the usual process taken, but the best behavioral studies aim to reach explanatory and predictive answers regardless of the initial understanding of behavior. .................................................................................................19 Figure 2.1: Acoustic telemetry activities at the Hammond Bay field site. A schematic of the VPS array that was located in Lake Huron around the mouth of the Ocqueoc River (blue line). Triangles represent receiver (VR2W) positions. VPS array testing in 2010 included two stationary tag tests (Gray dots, with median point as a black dot) and three mobile test transects (black dots forming lines). The schematic is oriented with north up and the black line running from left to right (east to west) represents the coast. .....................................................................49 Figure 2.2: The stationary test schematics and 2DRMS plots of all stationary test positions. Two schematics depict all VPS positions during two stationary tag tests, including (a) 29,355 positions between the dates (6/17/2010 to 7/01/2010), and (b) 16,400 positions between the dates (7/01/2010 to 7/08/2010), allowing us to evaluate array performance at two locations through an extended period of time. The white dot in the center of the clusters is the median location. The HPE versus measured error to the median point is shown for each estimated position during test one (c) and test two (d). The white circles with black outline and red x represent twice the distance root mean square error of x and y components of error within an HPE bin of one; 95% of tag detections have an error less than this point within each bin. Note there is a minimum HPE of 2.7 and 2.5 within the data. The line running between these points represents the 2DRMS and the equation and fit for this line are shown in the top left corner of (c) and (d), respectively. Data points above the 15 m bin, which can be seen in (a) and (b) are not shown in (c) or (d), because they are outside of the zone of interest. .........................................................................................50 Figure 2.3: Proportion of test positions with measured error from 1 to 10 m. Percent of positions with accuracy equal to or less than each measured value for the mobile tag test (circle), stationary tag at location one (triangle), and the stationary tag at location two (square) depicted an array with most positions having accuracy better than 6 m. The average error of the unfiltered data for the stationary tag at location one (1.98 m), location two (1.11 m), and the mobile test (6.83 m) are marked by representative symbols along the x-axis. ................................................52 Figure 2.4: Resultant data quality for HPE cutoffs of 3 to 15. The mean and maximum measured error is below 1.77 m for all HPE thresholds, sufficient to meet criteria 1 for both stationary test one (black) and test two (red). The maximum error exceeds 15 m at an HPE of 7 for test one and 8 for test two (violating criteria 2). The number of positions violating each test is located at the top of the figure for HPE cutoffs of 3 to 15. .................................................................................53 xiii Figure 2.5: Data loss versus error retention for HPE cutoffs of 3 to 15. The relationship of the percent of incorrectly rejected positions of all acceptable positions and percent of incorrectly retained positions of all retained positions suggested that HPE cutoffs of 8 to 10 for stationary test one and 3 to 15 for stationary test two, met criteria 4. ...........................................................54 Figure 2.6: An evaluation of the performance of the VPS array with a mobile tag. (a) A schematic depicting receiver positions (+) and the coast (black line) during the 2010 research season. A mobile tag test was completed on 7/6/2010. The small dots represent the VPS estimated positions during the mobile test. There were a total of 126 correctly retained positions (black , <6 m error, 15 m error) were enumerated for all three tests (two stationary and one mobile). HPE evaluation with test data The average error (criteria 1) and number of positions with large error were estimated (>15 m, criteria 2) for all HPE cutoff possibilities (3 to 15) to evaluate the first two criteria (Table 2.1). To evaluate if the HPE value provided for fish tags and calculated by VEMCO using synch-tag detections was a representative estimate of locational error for fish tags, we calculated 2DRMS for each stationary test tag by first calculating the Euclidian distance between each individual test position and the median or best estimate of the ‘true’ test tag location. The HPE 32 calculated by VEMCO is scaled to an approximately 1:1 relationship between HPE and measured error in m for synch-tags [9]. 2DRMS can be used to compare the relationship to fish tags and is reported in the base line evaluation of the dataset. The 2DRMS for the median value is actually a measure of precision [8], but it was also the best available estimate of accuracy. The 2DRMS linear model was calculated by first binning all data by one unit HPE increments and an average of the HPE within the bin was calculated to represent the bin. The error in the x direction (Xe) and y direction (Ye) was estimated for each location within a bin. The 2DRMS was then estimated for the 95% confidence interval using Xe and Ye within each bin and a line was fit to  the 2DRMS data 2DRMS Error = 2* SD(Xe) 2  + SD(Ye) 2  . A linear model fit to those 2DRMS values for each HPE (2DRMS line) of interest (3 to 15) was used to predict the proper HPE filter for 95% confidence about a given target error value. The 2DRMS regression developed for the mobile test was not considered due to the small number of recorded transmissions (138). The 2DRMS model was used to estimate the HPE at which there was 95% confidence that positions had less than 6 m accuracy (criteria 3). A range of HPE cutoffs (3 to 15) for removing unacceptably erroneous positions (>6 m error) while avoiding the loss of acceptable positions (<6 m error) were considered based on the percentage of incorrectly retained (of all retained) and incorrectly rejected positions (percentage of acceptable positions) (Table 2.2). The goal for the incorrectly retained percentage of all positions was >99% and incorrectly rejected of all positions was >95%, as the loss of data was considered more acceptable than allowing the erroneous positions to remain. This was an arbitrary proportion, though the loss of positions when evaluating trajectory data can quickly become problematic and retaining unacceptably erroneous positions could result in incorrect habitat assignment, so we 33 selected high target proportions. The variation in temporal gaps was not considered but could be very important to a study. Filtering fish data We examined all fish data classified by HPE in a schematic to assess any spatial pattern in HPE and to identify areas of poor array coverage. For each candidate HPE cutoff, we calculated the amount of data rejected inside and outside of the array area as defined as a polygon drawn around the outer most receivers. The post-filtering spatial coverage of the array was also tabulated for all candidate HPE values with the expectation that locations outside of the array are likely to have an increased HPE and lower accuracy [12] with increasing distance from the array periphery. Coverage was calculated by drawing a polygon around the outer most fish positions retained after filtering, and calculating the area (km2). Lastly, the evenness of filtered positions, referring to how grouped filtered positions were in the array, was considered for each individual fish to determine if specific subjects were more prone to being rejected. We visually assessed whether these positions were associated with any particular four-receiver diamond or if they were missed throughout the array, occurring in multiple four-receiver diamonds. RESULTS We demonstrate our approach by filtering the example dataset according to the data quality objectives of the study. Step 1: Establishing data quality objectives We adopted four data quality objectives. Data quality objectives are explicit assignments of acceptable errors in precision that derive from the requirements of the data analysis technique. 34 Any defensible objective could have been adopted to fulfill this step because the details are study-specific and could be simplified to a single criterion. Our four objectives included the ability to detect changes in trajectories (e.g., ground speed, turn angle, etc.), perform basic behavioral assignments (e.g., is the animal moving or stationary?), assign fish positions to habitat types (point data), and avoid the loss of acceptably accurate data while ensuring the removal of unacceptable positions if filtering is necessary to perform habitat assignment (Table 2.1). These objectives include both position specific criteria (i.e., <15 m error) and global criteria (i.e., mean error <1.77 m), though in each case a position specific filter (i.e., precision target) must be identified. The filter value should represent the minimum value that must have been met by a filtered dataset in order to achieve each objective. Among candidate filters, we identified the preferred filter as the one with the largest HPE that best met all criteria. The criteria provided a clear framework with which to select the best filter. Step 2: Estimating the baseline position confidence in the array Baseline position confidence refers to the quality of the unfiltered position data and is partly a function of array coverage (i.e., low quality positions are likely more abundant in regions of the array with poor coverage). Baseline confidence must be identified to determine if filtering is required. The baseline position confidence was described by mean and median array accuracy; position yield, the proportion of data that was below 6 m error, and the number of positions with greater than 15 m error were enumerated for stationary and mobile test tags that collectively determined if the array met the first three criteria. The fourth criteria dealt with data reduction. The initial evaluation was based on position yield (percentage of positions estimated vs. expected during the time period), average data accuracy, presence of large errors (>15 m), the twice the 35 distance root mean square (2DRMS) model equations, the proportion of data that was below 6 m error, and array coverage based on actual fish telemetry data. There were three tests of accuracy and each test varied in length, including 29,355, 16,400, and 138 positions (stationary site one, two, and mobile test) (Figure 2.1). The position yield for the stationary tags were 82.3% (location #1) and 89.4% (location #2), and 94.8% for the mobile test; this indicates that 5.2 to 17.7% of the tag transmissions did not result in a position. The estimated positions of the stationary tags were generally clustered around a central position; each exhibited a ‘tail’ of increasingly erroneous positions offset to the west. Without filtering, the mean error was 11.7 m for stationary test one (median: 2.9 m, range: 2.7 to 29,289.3 m), 4.2 m for stationary test two (median: 2.6 m, range: 2.5 to 1,425.8 m), and 5.81 m for the mobile test (median: 2.95 m, range: 2.50 to 186.40 m). There were 625, 134, and 0 positions with greater than 15 m of error (stationary test one, stationary test two, mobile test). The linear models for the 2DRMS regression line relating HPE and measured error obtained from the stationary-location tags were y = 0.18x – 0.89 (r2 = 0.92) for location #1 and y = 1.1x – 0.57 (r2 = 0.78) for location #2 (Figure 2.2). The slopes were both near 1 (0.81 and 1.1), indicating a similar relationship of HPE and synchronization tags to HPE and fish tags, with HPE potentially representing a conservative estimate of position error for fish tags (the slope of the 2DRMS line was less than 1). The proportion of test positions with 6 meter accuracy exceeded 90% for all three tests (Figure 2.3). Maximum array coverage was 8.22 km2, estimated by forming a polygon around the outer most estimated fish positions. Step 3a: Evaluating HPE filters consistent with the data quality objectives and guides The necessary filter cutoff for each criterion to be met must be determined prior to selection of a filter cutoff. Although the highest acceptable HPE value will retain the most data, 36 the range of HPE values that fulfill each criterion to reach each objective should be evaluated and reported. If no value meets all criteria, either the analysis associated with an objective must be reevaluated, a different analysis should be selected, or the research question will prove difficult to address with the original planned approach. Criterion 1 Any HPE cutoff from 3 (minimum observed in the test data) to 15 (maximum considerable due to objective 2) was sufficient to meet criterion 1 for stationary tests but not for the mobile test, although the mobile test methodology was susceptible to inherently high average error (Table 2.2). Criterion 2 The number of positions that violated criterion 2 increased greatly between HPE filters of 6 and 15 (max error: location 1: 10.0 to 26.7 m, location 2: 9.2 to 25.0 m; violating positions: location 1: 0 to 21, location 2: 0 to 10) (Figure 2.4). Only HPE cutoffs less than 7 (stationary location 1) or 8 (stationary location 2) met criteria 2, although no position had greater than 15 m error for the mobile test (Table 2.2). Criterion 3 The 2DRMS equation for desired 95% confidence in 6 m accuracy (Criterion 3) returned an HPE cutoff of 8.5 for stationary test one and 6.0 for stationary test two (Figure 2.2). There was not enough mobile test data to calculate this metric. 37 Criterion 4 The incorrectly retained proportion of all positions was less than 1% for HPE cutoffs of 3 to 10 for stationary test one and was 1% for all HPE cutoffs 3 to 15 for stationary test two (Figure 2.5). For the mobile test, only an HPE of 3 reduced the incorrectly retained proportion below though only four positions were incorrectly retained (3% of total; max. error = 7.7 m) when HPE cutoffs ranged from 6 to 10 (Figure 2.6; Table 2.2). Incorrectly rejected positions occurred at less than 5% of all positions for any HPE cutoff greater than 8 for stationary test one, greater than 3 for stationary test two, and greater than 5 for the mobile test, representing a minimal loss of acceptable data. When comparing the relationship between incorrectly retained and incorrectly rejected, acceptable HPE cutoffs of 8 to 10 (stationary test one), 4 to 15 (stationary test two), and no HPE was effective for the mobile test, although 5 to 10 was closest. Step 3b: Selecting an HPE cutoff The highest HPE cutoff that met or was the closest to meeting all criteria was an HPE of 8, although from this dataset HPE cutoffs between 6 to 8 all fit the criteria similarly. There were only two positions remaining with error greater than 15 m and an HPE less than 8 (criterion 2) during the combined 45,744 transmissions during the two stationary tests, which may be good enough for first passage time analysis if visual inspection of the remaining data points allows the remaining problematic values to be easily identified or the analysis for this step may need to be adapted. If we selected a lower HPE to meet this criterion perfectly, an HPE cutoff of 6 would have been required. An HPE cutoff of 6 would have incorrectly rejected 7.7% of data from stationary test one. Although a practitioner could make a case for this level of filtering, we preferred the increased coverage as data loss is a major issue with trajectory based analyses (Figure 2.7); 4,124 more positions would have been lost with an HPE of 6 vs. 8 (28.35% lost 38 outside array and 11.22% lost inside array for HPE 6) and total areal coverage would have reduced from 2.11 km2 (HPE 8) to 1.89 km2 (HPE 6) (Figure 2.8). Step 3c: Evaluating the potential for introduced bias through application of HPE cutoff There was evidence for spatial bias in filtering inside versus outside of the array and potentially behavioral or habitat filter bias. Only three mobile test positions with acceptable accuracy (<6 m) were rejected and all occurred at the array edge, suggesting that HPE may overfilter at the periphery of coverage (Figure 2.6a). In the sea lamprey data set, the HPE cutoff (8) rejected 8.8% of 58,025 total positions inside the array and 23.5% of 53,435 total positions outside of the array, though this comparison does not discriminate between removal of inaccurate and accurate positions (Figure 2.8). There was no evidence for spatial filter bias inside of the array as no region appeared more prone to poor positioning across individuals (Figure 2.8). A behavioral or habitat filter bias was evident, as the majority of positions rejected from inside of the array (3,070 positions, 74.6% of all rejections) were associated with seven lampreys during daylight hours that were likely stationary, as sea lamprey are nocturnal. This observation is consistent with previous observations that sea lamprey may settle in locations that interfere with acoustic tag signal transmission, blocking the line of sight between receivers and tags [19]. DISCUSSION We developed a straightforward conceptual approach for using a PPE to filter hyperbolically positioned data and demonstrated this approach with a technology (underwater acoustic telemetry) in which users evaluate position accuracy with HPE. The framework included selection of defensible data quality objectives, evaluation of the array’s positioning 39 accuracy with an independent dataset, and determination of the relationship of the selected PPE to measured accuracy and data retention. Data-filtering with a PPE estimate of position accuracy has certain conceptual advantages over a biological filter if the PPE is properly evaluated. PPE’s are calculated for each position obtained from the telemetry apparatus, whereas biological filters ignore the positioning process and only evaluate resultant positions based on an expectation of what is biologically reasonable for the study species. Biological filters can be useful and are frequently used because they are conceptually straightforward, and at times, the only available option. Some habitat filters are quite reasonable (e.g., fish do not swim 500 m onto land); although the rule could introduce filter bias as positions closer to the physical habitat edge are more likely to be rejected. Similarly, as HPE represents a 95% confidence value, HPE becomes large when solutions become less precise, as typically occurs outside the array periphery where the overlapping of parabolas allows for multiple potential solutions [9]. Failure to select a proper filter is more problematic for calculating movement trajectories (e.g., maximum ground speed), and identifying a useful biological filter is especially challenging for aquatic species for which maximum movement capacities are often unknown [22], or poorly estimated. Selection of a biological filter cutoff that is high allows incorrectly retained positions to remain in the dataset, whereas cutoffs that are low near the average speed remove valuable data. Either case will serve to infuse the data with a perceived improvement in accuracy that is not supportable (i.e., the rejected positions may be no less accurate than many retained positions). Unlike biological filters, PPE filters are position specific and rely principally on the assumption that animal-integrated tags match the performance of stationary or towed tags, and that the array is well-constructed to ensure sufficient areal coverage, avoidance of obstructions, etc. [9,23]. If these assumptions are supported, carefully selected PPEs that are based on the position quality should be used prior to 40 biological filters that are based upon the biological plausibility of resultant positions and not the quality of the positioning process. PPE filter selection ranges from choosing an arbitrary HPE based on subjective operator preferences to developing a complete algorithm that would output a filter threshold based on a set of inputs (i.e., the model determines both performance and value). The approach we suggest clearly falls in the middle (i.e., performance measurement is objective but value judgment can be subjective) but still represents a significant improvement over the use of arbitrarily selected filter values. The selection of an HPE of 8 fell within the most effective range for the criteria and ensures high confidence in 6 m accuracy (criteria 3). Although we identified three different criteria to complete a complex analysis, a single criterion could be chosen for a single analysis, or, if multiple analyses are contemplated, a different cutoff could be chosen for each, which would make selection and reporting straightforward. In our example, the criteria for objective 2 was not met, as it required no positions to have greater than 15 m of error, clearly representing an ineffective criteria for an objective. At very low numbers of violating positions, only extreme positions remain and improving the filter to remove these positions came at a high cost (increase in positions incorrectly rejected). These extreme outliers may not even represent predictable performance of the system. If the filter is to be useful, the proper response is to adjust the first passage time analysis, which is very easy and would have only required shifting the moving designation by a few meters (17.7 m was the largest remaining error). However, perhaps a better criterion would be to choose a high percentile (e.g., 99.5%) that represents an acceptable level of risk or error in your future analysis. With the VPS system, little effort has been made to defend filtering cutoffs beyond reference to prior use [13], and ambiguous filter criteria are at risk of inadvertently becoming acceptable practice through the accumulation of use. In our case, adoption of an ambiguous filter based on a previous study (HPE 10 to 20 [14-16]), would have 41 been less useful than the carefully evaluated filter cutoff of 8 and indefensible (Table 2.2). Telemetry technology represents a very different tool than typical scientific instruments as its design is rarely consistent, is difficult to standardize, and does not generate data points with fixed accuracy and precision. For this reason, we suggest that data filtering should be a flexible process that may progress towards more concrete rules if some level of standardization of use occurs for telemetry equipment as has occurred for the use of positional dilution of precision in certain GPS applications (e.g., [24]). A VPS system is capable of attaining accurate positions (2DRMS <6 m, <2 m average) with a high position yield (>82%) via autonomous receivers that are capable of covering large areas (>2 km2), although these results were specific to this system and environment. As with all systems, the VPS was susceptible to spatial, temporal, and behavioral or habitat bias in position yield and position precision which could cloak important biological phenomena [6,24]. For example, when an animal occupies a habitat that blocks the line of sight from tags to receivers, as was suspected for the seven stationary sea lamprey that composed 74.6% of our filtered data inside of the array [19], there is a potential habitat and behavioral bias [9]. Observed HPE values also increased with distance outside of the test array, consistent with observations from other studies [9,13]. Confirming a temporal or spatial bias is challenging because, depending on placement, stationary tags may not reveal systematic spatial biases in the array [9,25], and may be sensitive to regular variation in environmental conditions (e.g., louder waves in shallow vs. deep waters) that differentially impact receiver performance. Mobile range testing is recommended for spatial evaluations of filter performance, though results may not be representative of the full range of environmental conditions encountered by tagged animals due to the typically short duration of mobile tests, selection of favorable boating conditions, etc. The 42 mobile test appeared to be biased towards higher error estimates, though the spatial patterns were consistent between fish tracking and mobile test data. Both mobile and stationary tests presented unforeseen challenges and unexpected findings. Mobile testing suggested an HPE of 8 was overly conservative outside of the internal array area, though we lacked stationary tests in this region that would have confirmed this conclusion (Figure 2.6). The mobile test presented some challenges to evaluation because the GPS clock was not synchronized with VPS time and only recorded a position every second. To minimize the effects of clock differences on position error estimates, we applied a constant time offset to all mobile test positions that minimized error between the mobile test tag tracks and corresponding GPS tracks (Additional file 1). In addition, we likely over-estimated error by assuming that the GPS track represented the true path of the test tag because we only collected a single post-processed position every second along the track. Collecting positions with sub-meter accuracy usually requires averaging several GPS points at a given location, but point averaging over time is not an option while moving. The tail of positions that were observed in the stationary tests could not be explained with the example dataset but was likely the result of a specific set of receivers with poor geometry that consistently estimated inaccurate positions in one direction. The tails were not troubling to us as these positions were easily filtered with HPE. Many challenges can be avoided in advance by careful project planning. We recommend multiple mobile and stationary tag tests with the animal tag during the same time period that the animals are being monitored, and ideally, at least some fraction of the synch-tags would be the same model as the fish tags. Receivers should be positioned outside the maximum spatial extent of interest, or at a minimum, stationary animal tags should be monitored in any area of interest to ensure the chosen HPE cutoff has met the filter goals. 43 Although we present a process for selecting a single fixed HPE cutoff and provide the R code necessary for performing this approach (Additional file 1), a dynamic filter in which the HPE cutoff could be either temporally (e.g., [26]), or both temporally and spatially flexible, might prove useful, although it also would require synchronization tags that match fish tags. Regardless of how the HPE selection process is fine-tuned, the key is following a standard process like the one we have described, including reporting the process for others to properly assess the research findings. Even if the necessary accuracy required to perform the mostpreferred analysis cannot be attained with the data collected, PPEs can be used to tailor the selection of an analytical tool, or the spatial scale at which the behavior is considered. Either adaptation to the study is preferred over the reporting of errant observations [17,27]. Conclusions PPE error estimates are frequently available from animal telemetry systems that rely on hyperbolic positioning and can be used to evaluate data quality prior to analysis. When using PPE to filter data, practitioners should undertake (a) a priori determination of data accuracy requirements; (b) independent assessment of the telemetry system performance; (c) a determination of how well the PPE represents measured accuracy; (d) selection of a filter cutoff based on the balance between accuracy improvement and data retention; and (e) explicit consideration of spatial, behavioral, and habitat bias associated with the telemetry system and the animal under observation. A carefully constructed PPE filter is more defensible than biological filters that can improve data accuracy but require (1) an interpretation of the data vs. an assessment of its precision, and (2) are only applied to a subset of the data collected (extreme movements). HPE offers the intriguing possibility for direct use in the analysis as an error estimate (vs. a criterion for data retention); akin to bench apparatus precision estimates, though 44 there is no evidence that this approach has been used in such a manner with other hyperbolic positioning systems. Because data analysis requirements are likely to be as varied as the movement data to which they are applied, complete exposition of the selection process and criteria should be included in the methods section of any subsequent reports or publications. The minimum level of a reporting should include a description of the data quality objectives, criteria, rational for cutoff selection, and evidence of reaching the criteria, which could come in the form of a paragraph, table, or appendix and does not need to be a substantial component of the report or paper. ACKNOWLEDGEMENTS This manuscript is contribution number 1843 of the Great Lakes Science Center. This work was funded by the Great Lakes Fishery Commission by way of Great Lakes Restoration Initiative appropriations (GL-00E23010-3). This paper is contribution 7 of the Great Lakes Acoustic Telemetry Observation System (GLATOS). 45 APPENDIX 46 Table 2.1: Filtering Objectives. The four specific criteria adopted to establish the data quality objectives for the project. Objective Criteria Rationale 1 Detect changes in trajectory Mean error ≤1.77 m Many trajectory based analyses are only preserved when the average position error is <10% of the mean step length [17]. Transmission delay was (33.4 s), sea lamprey ground speed is 0.53 ms−1 [18,19], equating to a step length of 17.7 m. 2 Assign behavioral state Max error <15 m (moving or stationary) to fish positions 3 Assign habitat to fish positions 2DRMS ≤6 m To ensure accurate assignment of each estimated fish position to habitat types that are defined at the resolution of 18 × 18 m grid cells we selected a twice the distance root mean square (2DRMS) error of <6 m for our selected HPE values. 4 Balance loss of acceptable data and retention of unacceptable data Retain 95% of acceptable data, Reject ≥99% of unacceptable data To ensure accurate assignment of each estimated fish position to habitat types while avoiding loss of accurate data. We aimed to keep 95% of acceptable data (≤6 m) while retaining <1% of unacceptable data (>6 m) of all retained positions (unacceptable data remaining/all retained positions × 100). Simple behavioral assignment of moving or stationary with first passage time analysis, which is a measure of the time it takes an individual to leave a circle of fixed radius r drawn around each measured location to determine if movement is occurring [20,21]. Based on prior analyses, a displacement of greater than 15 m per transmission at the average tag transmission rate (33.4 s) could result in a false designation of moving when the fish was stationary. 47 Table 2.2: Criteria for selection of an HPE filter cutoff Criteria Stationary Mobile test test 1 2 3 This criterion was met for stationary tests (Figure 2.4), but the mobile test would have required a very low HPE to attain below 1.77 m of average error as the unfiltered accuracy was lower (Figure 2.3). See review of the mobile test methodology in the Discussion. ≤15 An HPE of 8 did not meet the criteria for stationary tests, but only 1 of 625 positions (test one) and 1 of 134 positions (test two) remained. The mobile test criteria were met. Only 2 of 45,744 positions were problematic (<0.001%) for the combined stationary tests (Figure 2.4). The maximum point remaining was 17.7 m error. NA The criterion was met for stationary test one, which was the longer test that covered more variable weather conditions. The HPE 8 bin only was 95% confident in an HPE of 8.25 meters for test two but the estimate was based on data calculated within each bin and few points had an HPE of >6 for test two (3% of the data), which results in a less reliable 2DRMS prediction. Criterion 1 3–15 3–15 Mean ≤1.77 m Criterion 2 <7 <8 Max error <15 m Criterion 3 2DRMS ≤6 m 8.5 6.0 Rational for selection of a HPE filter cutoff of 8 Criterion 4 None Percentage incorrectly 8–10 3–15 retained vs. percentage incorrectly rejected 5–10 An HPE cutoff of 8 met the criteria for both stationary tests, and although the mobile test did not have a suitable range, the range from 5–10 was equally effective. 48 Figure 2.1: Acoustic telemetry activities at the Hammond Bay field site. A schematic of the VPS array that was located in Lake Huron around the mouth of the Ocqueoc River (blue line). Triangles represent receiver (VR2W) positions. VPS array testing in 2010 included two stationary tag tests (Gray dots, with median point as a black dot) and three mobile test transects (black dots forming lines). The schematic is oriented with north up and the black line running from left to right (east to west) represents the coast. 49 Figure 2.2: The stationary test schematics and 2DRMS plots of all stationary test positions. Two schematics depict all VPS positions during two stationary tag tests, including (a) 29,355 50 Figure 2.2: (cont’d) positions between the dates (6/17/2010 to 7/01/2010), and (b) 16,400 positions between the dates (7/01/2010 to 7/08/2010), allowing us to evaluate array performance at two locations through an extended period of time. The white dot in the center of the clusters is the median location. The HPE versus measured error to the median point is shown for each estimated position during test one (c) and test two (d). The white circles with black outline and red x represent twice the distance root mean square error of x and y components of error within an HPE bin of one; 95% of tag detections have an error less than this point within each bin. Note there is a minimum HPE of 2.7 and 2.5 within the data. The line running between these points represents the 2DRMS and the equation and fit for this line are shown in the top left corner of (c) and (d), respectively. Data points above the 15 m bin, which can be seen in (a) and (b) are not shown in (c) or (d), because they are outside of the zone of interest. 51 Figure 2.3: Proportion of test positions with measured error from 1 to 10 m. Percent of positions with accuracy equal to or less than each measured value for the mobile tag test (circle), stationary tag at location one (triangle), and the stationary tag at location two (square) depicted an array with most positions having accuracy better than 6 m. The average error of the unfiltered data for the stationary tag at location one (1.98 m), location two (1.11 m), and the mobile test (6.83 m) are marked by representative symbols along the x-axis. 52 Figure 2.4: Resultant data quality for HPE cutoffs of 3 to 15. The mean and maximum measured error is below 1.77 m for all HPE thresholds, sufficient to meet criteria 1 for both stationary test one (black) and test two (red). The maximum error exceeds 15 m at an HPE of 7 for test one and 8 for test two (violating criteria 2). The number of positions violating each test is located at the top of the figure for HPE cutoffs of 3 to 15. 53 Figure 2.5: Data loss versus error retention for HPE cutoffs of 3 to 15. The relationship of the percent of incorrectly rejected positions of all acceptable positions and percent of incorrectly retained positions of all retained positions suggested that HPE cutoffs of 8 to 10 for stationary test one and 3 to 15 for stationary test two, met criteria 4. 54 Figure 2.6: An evaluation of the performance of the VPS array with a mobile tag. (a) A schematic depicting receiver positions (+) and the coast (black line) during the 2010 research 55 Figure 2.6: (cont’d) season. A mobile tag test was completed on 7/6/2010. The small dots represent the VPS estimated positions during the mobile test. There were a total of 126 correctly retained positions (black , <6 m error, 8) Figure 2.8: Fish positions filter by an HPE cutoff of 8. A depiction of how many of all fish locations in 2010 would be rejected by an HPE filter of 8.0. Fish locations are represented by small black dots, receiver positions are shown by red triangles and the coast is depicted by a gray line. These are cumulative graphs with all sea lamprey tags shown in (a), fish positions with an HPE below or equal to 8.0 are displayed in (b), and fish positions with an HPE greater than 8.0 are shown in (c). 58 Figure 2.8: (cont’d) Supplementary Material: R code for filtering VEMCO positioning system data with Horizontal Positioning Error (HPE). Available at: http://www.animalbiotelemetry.com/content/2/1/7/additional 59 REFERENCES 60 REFERENCES 1. Cagnacci F, Boitani L, Powell RA, Boyce MS: Animal ecology meets GPS-based radiotelemetry: a perfect storm of opportunities and challenges. Philos Trans R Soc London [Biol] 2010, 365(1550):2157–2162. 2. Cooke SJ, Hinch SG, Wikelski M, Andrews RD, Kuchel LJ, Wolcott TG, Butler PJ: Biotelemetry: a mechanistic approach to ecology. Trends Ecol Evolut 2004, 19(6):334–343. 3. Cooke SJ, Midwood JD, Thiem JD, Klimley P, Lucas MC, Thorstad EB, Eiler J, Holbrook C, Ebner BC: Tracking animals in freshwater with electronic tags: past, present and future. Anim Biotelem 2013, 1:5. 4. Giuggioli L, Bartumeus F: Animal movement, search strategies and behavioural ecology: a cross‐disciplinary way forward. J Anim Ecol 2010, 79(4):906–909. 5. D’Eon RG: Effects of a stationary GPS fix rate bias on habitat selection analyses. J Wildlife Manage 2003, 67:858–863. 6. D'Eon RG, Delparte D: Effects of radio‐collar position and orientation on GPS radio‐ collar performance, and the implications of PDOP in data screening. J Appl Ecol 2005, 42(2):383–388. 7. Frair JL, Nielsen SE, Merill EH, Lele SR, Boyce MS, Munro RHM, Stenhouse GB, Beyer HL: Removing GPS collar bias in habitat selection studies. J Appl Ecol 2004, 41:201–212. 8. Whithey JC, Bloxton TD, Marzluff JM: Effects of tagging and location error in wildlife telemetry studies. In Radio Tracking and Animal Populations. Millspaugh JJ, Marzluff JM (Eds.). Waltham, MA: Academic Press Inc.; 2001, pp. 43–75. 9. Smith F: Understanding HPE in the VPS Telemetry System. VEMCO Tutorials; 2013. [http://vemco.com/wp-content/uploads/2013/09/understanding-hpe-vps.pdf] 10. Andrews KS, Tolimieri N, Williams GD, Samhouri JF, Harvey CJ, Levin PS: Comparison of fine-scale acoustic monitoring systems using home range size of a demersal fish. Mar Biol 2011, 158(10):2377–2387. 11. White GC, Garrott RA: Analysis of Wildlife Radio-Tracking Data. Waltham, MA: Academic Press Inc.; 1990. 12. Espinoza M, Farrugia TJ, Webber DM, Smith F, Lowe CG: Testing a new acoustic telemetry technique to quantify long-term, fine-scale movements of aquatic animals. Fish Res 2011, 108(2):364–371. 61 13. Roy R, Beguin J, Argillier C, Tissot L, Smith F, Smedbol S, De-Oliveira E: Testing the VEMCO Positioning System: spatial distribution of the probability of location and the positioning error in a reservoir. Anim Biotelm 2014, 2:1. 14. Scheel D, Bisson L: Movement patterns of giant Pacific octopuses, Enteroctopus dofleini (Wülker, 1910). J Exp Mar Biol Ecol 2012, 416–417:21–31. 15. Furey NB, Dance MA, Rooker JR: Fine‐scale movements and habitat use of juvenile southern flounder Paralichthys lethostigma in an estuarine seascape. J Fish Biol 2013, 82:1469–1483. 16. McMahan MD, Brady DC, Cowan DF, Grabowski JH, Sherwood GD: Using acoustic telemetry to observe the effects of a groundfish predator (Atlantic Cod, Gadus morhua) on movement of the American lobster (Homarus americanus). Can J Fish Aquat Sci 2013, 70(11):1625–1634. 17. Bradshaw CJA, Sims DW, Hays GC: Measurement error causes scale-dependent threshold erosion of biological signals in animal movement data. Ecol Appl 2007, 17(2):628– 638. 18. Vrieze LA, Bergstedt RA, Sorensen PW: Olfactory-mediated stream-finding behavior of migratory adult sea lamprey (Petromyzon marinues). Can J Fish Aquat Sci 2011, 68:523–533. 19. Meckley TD, Wagner CM, Gurarie E: Coastal movements of migrating sea lamprey (Petromyzon marinus) in response to a partial pheromone added to river water: implications for management of invasive populations. Can J Fish Aquat Sci 2014, 71(4):533– 544. 20. Fauchald P, Tveraa T: Using first-passage time in the analysis of area-restricted search and habitat selection. Ecology 2003, 84(2):282–288. 21. Barraquand F, Benhamou S: Animal movements in heterogeneous landscapes: identifying profitable places and homogenous movement bouts. Ecology 2008, 89:3336–3348. 22. Katopodis C, Gervais R: Ecohydraulic analysis of fish fatigue data. River Res Applic 2012, 28:444–456. 23. Espinoza M, Farrugia TJ, Lowe CG: Habitat use, movements and site fidelity of the gray smooth-hound shark (Mustelus californicus, Gill 1863) in a newly restored southern California estuary. J Exp Mar Biol Ecol 2011, 401(1):63–74. 24. Lewis JS, Rachlow JL, Garton EO, Vierling LA: Effects of habitat on GPS collar performance: using data screening to reduce location error. J Appl Ecol 2007, 44(3):663– 671. 62 25. Biesinger Z, Bolker BM, Marcinek D, Grothues TM, Dobarro JA, Lindberg WJ: Testing an autonomous acoustic telemetry positioning system for fine-scale space use in marine animals. J Exp Mar Biol Ecol 2013, 448:46–56. 26. Coates JH, Hovel KA, Butler JL, Klimley AP, Morgan SG: Movement and home range of pink abalone Haliotis corrugate: implications for restoration and population recovery. Mar Ecol Prog Ser 2013, 486:189–201. 27. Hayes GC, Kesson SA, Godley BJ, Luschi P, Santidrian P: The implications of location accuracy for the interpretation of satellite-tracking data. Anim Behav 2001, 61:1035–1040. 28. Nocedal J, Wright SJ: Numerical Optimization. New York, NY: Springer; 1999. 63 CHAPTER 3 HOW DO NON-HOMING FISHES FIND THE SHORE? EVIDENCE FOR BATHYMETRIC ORIENTATION IN MIGRATING SEA LAMPREY ABSTRACT How non-homing fishes like the sea lamprey (Petromyzon marinus) find distant targets is poorly understood. Following translocation by host fishes during parasitic feeding in the Great lakes, sea lamprey must find the coast and enter a river to spawn. Three-dimensional paths of 22 captured and released spawning-phase sea lamprey were quantified from the commencement of movement from a settled state until exiting an acoustic array with 3 km2 of coverage, centered 3.3 km from the western coast of Lake Huron in greater than 33 m of water. 81 % of individuals arrived at the nearest coast within 72 hours, suggesting efficient orientation to the coast termed y-axis orientation. Two phases of movement were documented for most individuals: i) in phase 1, sea lamprey moved slowly on the bottom and turned in a consistent direction and made vertical ascents to the surface and returned to the bottom; and, ii) in phase 2, sea lamprey moved more quickly and steered a straight course while making occasional heading corrections and pirouettes and moved primarily at the surface, with occasional descents to the bottom. During phase 2, sea lamprey moved in the direction of the bathymetric slope towards shallower water rather than towards the nearest coast, and no apparent effect was observed due to where the sea lamprey had previously entered a river in Michigan. We postulate that these sea lamprey used the absolute hydrostatic pressure sensed at successive sites along the lake bottom to pick a heading towards shallower water during Phase 1 and used declining hydrostatic pressure at the lake bottom to steer towards shallower water during Phase 2. In contrast to a natal homing migration, 64 the non-homing sea lamprey migration is consistent with a growing number of observations of organisms that orient to a region of resource availability with a simple set of movement rules based on topographic features of the environment. INTRODUCTION Efficient navigation to a distant target requires that a searching animal must first orient itself to environmental features (referents) that reveal the direction to the target, undertake directed movements in accordance with the referent(s) (Able 1991, Åkesson and Hedenström 2007, Luschi 2013), and recognize the target upon arrival (Dittman and Quinn 1996, Walker et al. 2002). Animals can reach resources via: i) non-oriented search (random), ii) oriented search with information that is independent of the resource but related to a region of high target availability, iii) a cue directly associated to the resource, or iv) based on prior knowledge of the target location (Mueller and Fagan 2008). To understand how animals orient during migration, large scale detailed data on the environment, the movement path and an understanding of the internal state of the animal is necessary (Nathan et al. 2008). Studies of this nature in aquatic environments are rare (Lohmann et al. 2008, Cooke et al. 2013), as underwater positioning technologies and methodologies capable of large-scale coverage with high resolution are just becoming available (Jonsen et al. 2003, Espinoza et al. 2011, Biesinger et al. 2013). The emerging theory for homing migrations is that migrants switch between navigation methods at different scales but rely predominantly on geomagnetic information to arrive near their natal home before relying on other senses, notably olfaction, to move more specifically to a reproduction site (Lohmann et al. 2008, Putman et al. 2014). A major difference in non-homing oriented search is that individuals can rely on topography to locate a distinct geophysical region 65 with high resource availability (e.g., Pe'er et al. 2004). Orientation to topographic features is common to many species including hilltopping butterflies, Melitaea trivia, that move upslope to locate mates, sand hoppers, Orchestoidea corniculata, that rely on slope and other features to maintain perpendicular orientation to the shoreline in a preferred tidal zone (Craig 1973, Scapini et al. 1996), or Monarch butterflies, Danaus plexippus, that rely on simple geographic barriers to funnel their migration towards Mexican overwintering sites (Mouritsen et al. 2013). Geophysical features that inform an organism of its position are commonplace, though they may be masked by direct cues or not useful at all in a homing migration due to the specificity of their target. Specifically, many aquatic organisms orient themselves perpendicularly to the coast in a special case of orientation termed y-axis orientation (Ferguson 1971). Y-axis orientation has been observed in many species including multiple amphibians that rely on celestial features to orient to shore (Ferguson et al. 1965, Landreth and Ferguson 1967) or green turtles, Chelonia mydas, that use windblown odors to orient to a specific beach (Luschi et al. 2001). Over large distances in large bodies of water, moving opposite a gradient such as slope along the lake bottom could theoretically provide directional cues leading to a coast, which could be considered analogous to navigating up a hill. Orientation in a non-homing species is more difficult to study than in homing species because the animal has greater flexibility to use indirect information to orient that may not be directly associated to a specific target (Benhamou 2006). For this reason, orientation mechanisms have principally been studied in species with prior knowledge of the target, wherein referents can be manipulated to reveal if the change in the referent alters how the animal searches for a specific site (Lohmann et al. 2008, Holland et al. 2010, Luschi 2013). The time required for an animal to orient depends on how long it takes to gather directional information and calibrate an 66 internal compass. Such a situation often results in an initial period of sinuous exploratory movement (Wendler and Scharstein 1986, Ueda et al. 1998, Pe'er et al. 2004). When searching for distant targets, oriented animals usually move in a straighter fashion. The heading of the straight movement or direction of the final point observed in a study (vanishing point) can be compared with environmental features to determine if there is a consistent pattern and what information the animal is using to orient (Girard et al. 2004, Holland et al. 2010). The obligation of non-homing adult sea lamprey to efficiently find a coastline and subsequently enter a suitable river to spawn after feeding has terminated makes them an excellent candidate for studying y-axis orientation in fishes. The Great Lakes adult sea lamprey completes a non-homing reproductive migration from lacustrine feeding grounds, where they parasitize large bodied fishes, to riverine spawning habitat (Bergstedt and Seelye 1995, Waldman et al. 2008). The semelparous migration initiates when they release from their final host, which may occur at a location never previously visited, and may represent an uncontrolled displacement (Clemens et al. 2010, Silva et al. 2014). Any coast is a viable target for fishes in a lake (Meckley et al. 2014b), whose target is a river labeled locally by olfactory cues (Waldman et al. 2008, Vrieze et al. 2010). Therefore, sea lamprey could feasibly rely on indirect cues that allow for straight movement in any direction or cues directly associated to the nearest coast. Once arriving at the coast, sea lamprey move parallel to the coastal edge while searching for river water at the lake edge (Vrieze et al. 2011, Meckley et al. 2014b). If sufficient larvae occupy a river, water at the mouth will contain larval odor indicating adequate proven spawning and rearing habitat within the watershed (Teeter 1980, Sorensen et al. 2005, Wagner et al. 2006). Sea lamprey rely heavily on non-visual information during their nocturnal migration limiting potential orientation mechanisms (Binder and McDonald 2007, Keefer et al. 2013). Based on the 67 life history and sensory capacity of sea lamprey, we postulate that sea lamprey rely on an aspect of the local bathymetric gradient to determine the direction to shore (y-axis orientation). We investigated the three dimensional path of sea lamprey in Lake Huron by following their release 3.3 km from the nearest coast in 30 m of water to test whether they complete y-axis orientation and return to the coast by following the bathymetric gradient. We specifically evaluated a bundled set of predictions to address our complex hypothesis, (1) sea lamprey will appear at the nearest coastline within 72 hours, indicating that they performed y-axis orientation and approached the nearest coast first; (2) sea lamprey move in a two-phase pattern after displacement consistent with local exploration (high sinuosity) followed by search for a distance target (straight movement), indicating that we observed an orientation process; (3) the cardinal direction of movement during the second phase across individuals is not in a consistent direction, signifying a variable search strategy or use of a locally varying feature; (4) the cardinal direction will be consistent with following the local bathymetric gradient ; and (5) the direction will be based on local information in the array and not on which river the sea lamprey was trapped in Michigan. METHODS Ethics Statement The treatment and acoustic tagging of sea lamprey was approved by the Michigan State University Institutional Animal Use and Care Committee via animal use permit 02/10-020-00. Study Design The tracks of 22 sea lamprey were analyzed after their release in the center of an acoustic 68 positioning array located 3.3 km from the nearest coast (Figure 3.1). The array positioned fish over a region of the coast that had shallower bathymetric contours to the south and west depending on where movement commenced (Figure 3.2). Sea lamprey did not always stop quickly after release so movement could commence at different regions within the array despite release occurring at the center of the array. We bundled predictions to test our hypothesis that sea lamprey rely on an aspect of the bathymetric gradient to determine the direction to shore regardless of the compass direction of the bathymetric slope in front of the river in which they were trapped. To determine if sea lamprey performed y-axis orientation we relied on the first detection of each sea lamprey at acoustic receivers located along the coast (Figure 3.1). To determine if sea lamprey employed a two-phase strategy indicating that we captured the orientation process, we relied on visual inspection and evaluation of track sinuosity. The heading of the sea lamprey from the point it began moving in a straight path until exiting the array was extracted and used to determine if sea lamprey move directly towards the closest coast or follow the local bathymetric contours, and whether there was an effect of past experience, in terms of the river they were caught. Lastly we characterized the vertical movement of sea lamprey to build hypotheses why sea lamprey make excursions through the entire water column and what role it may have in search behavior. Typically sea lampreys were transported to the field site in aerated coolers and released in the center of the acoustic array at 15:00 when weather permitted (09:00 - 17:30 EST) by being lowered in a release cage at a rate of 6 m·min-1 that opened at a fixed depth of 30 m when door clamps were triggered by pressure. Three sea lamprey were released at the surface to determine the minimum rate at which lamprey could descend without apparent harm and all reached the 69 bottom in less than 2 min, equating to a descent rate of faster than 17.5 m·min-1. The surfacereleased sea lamprey were not considered in any other test. Sea lamprey came from six different sources, located in three different scenarios where the contours from deeper to shallower water (bathymetric slope) occurred in three different compass directions (Figure 3.1). If sea lamprey orient to any aspect of the coast and relied on geomagnetic information from previous experience we would expect these conditions to influence their selected headings. Adult female sea lamprey were obtained from five rivers in Michigan via barrier traps (Manistee River: 44.249981, -86.344531 ( N=8, trapped 15-May2012), Cheboygan River: 45.656202, -84.464478 (N=21, trapped 02-May-2012), Manistique River: 45.945189, -86.247733 (N=21, trapped 05-May-2012), Betsie River: 44.630058, 86.252273 (N=3, trapped 21-May-2012), Ocquoec River: 45.490246, -84.072981 (N=4)), and parasites were caught while attached to fish in the lake (Lake Huron, Hammond Bay (N=10, caught January 03,2012 February 28, 2012)) (Figure 3.1). Individuals from the Manistee and Betsie rivers did not move for several days after release and were dropped from the study. These animals were likely more mature and no longer migratory, as they came from rivers that were warmer. As a result, we compared only the Manistique and Cheboygan River sea lamprey to directly evaluate experience, in terms of the river in which they were trapped. Study Site The center of the study site (45.527799, -84.044466) was in Lake Huron and located 3.3 km from the nearest shoreline in an area of the lake where bathymetry varied only in two directions (Figure 3.1). A telemetry system (VPS, VEMCO, Halifax, NS) composed of 43 receivers (VR2W) in diamond formations with between receiver spacing of 275 m and an 70 internal area of 3 km2 recorded sea lamprey position (Figure 3.2; see Meckley et al. 2014a). Approval for temporary placement of equipment was provided by the United States Coast Guard (Permit: 16518, Ser. 09-12). Receivers were placed 3 m from the lake bottom in a region of the lake ranging from 30.6 m – 37.2 m deep, although some fish positions were obtained for fish located outside of the array in an area ranging from 21-39 m deep. Nine synchronization transmitters (VEMCO model V16-2H, 69 kHz, 160 db, 500 to 700 s transmission interval) were collocated with receivers to maintain time synchronization of the receiver clocks. In addition to the VPS array, individual receivers provided detection data along multiple locations on the coast and in the Black Mallard, Ocquoec and, St. Mary’s Rivers as part of the Great Lakes Acoustic Telemetry Observation System (GLATOS, Figure 3.1). Detections on these receivers informed us of whether sea lamprey reached the nearest coast prior to other coasts. Solitary receivers were positioned at the Black Mallard River Mouth (BM, 45.532888, -84.120801) and Ocqueoc River Mouth (OCQ, 45.491893, -84.071879). Receiver strings were treated as individual detection points and were located near the Cheboygan River mouth (CHB, 45.67239, -84.429368, N=5), 40 mile point (FMP, Outer receiver: 45.507563, -83.901379, N=3), Presque Isle (PRS, Outer receiver: 45.333842, -83.458343, N=3), and Detour Pass, which leads to the St. Mary’s River (SMR, West side of Drummond Island, centered at: 45.984929, -83.891787, N=6). Detection data revealed sea lamprey presence and was used to determine the fate of sea lamprey on the coast. Data were obtained from shared receivers as part of GLATOS. The detection range of a tag with 150 db of power placed in 5 m of water will vary and is based on line of sight to a receiver, and the environment. A 2-D range test in Hammond Bay under calm conditions revealed a 95% detection efficiency at 155 m and a 23% detection at 1.1 km. Detections exceeding 1.5 km and passing the nearest neighbor filter were unlikely. The detection 71 range of a receiver placed in < 3 m of water directly in front of the Ocqueoc and Black Mallard rivers, as was done in this study, was likely lower than observed in the range test, especially because the receiver was located in the wave zone. A nearest neighbor filter of 30 minutes was used to remove spurious detections and rare detections near the maximum extent from the GLATOS detection data. Bathymetry in the detection region of the array was measured along transects spaced 50 m apart and oriented northwest to southeast, and then repeated crisscrossing the first grid at an orientation of southwest to northeast. A Lowrance depth sounder (HDS-8) streamed depth and GPS location to a laptop to record bottom depth during calm conditions and was matched to more accurate post processed GPS positions collected by a Trimble GeoXH and Tornado antennae. Total water column depths at fish positions were interpolated in program R from the georeferenced depth data using inverse distance weighting (function “interp”, package akima) (Akima 1978, Gebhardt et al. 2013). Tagging Procedure All sea lamprey were held in 150 L flow-through tanks that cycled ambient Lake Huron water (100% exchange every 2 h) under a natural light cycle. Two types of acoustic transmitters transmitted signals every 15-45 s, though one contained a pressure sensor (n=49, model V9P2H,Vemco, Halifax, Nova Scotia, Canada 9mm D x 47 mm L, mass: 6.4 g in air, 3.5 g in water, power output 150 dB (re 1 µPa at 1 m)) and the other only provided horizontal position (n= 18, model V9-2H,Vemco, Halifax, Nova Scotia, Canada, 9mm diameter, 29 mm length: mass: 3.6 g in air, 2.2 g in water, power output 151 dB (re 1 µPa at 1 m)). Sea lamprey were 273-577 mm long (mean 499 mm) and weighed 139-398 g (mean 274 g). Acoustic tagging procedures 72 followed Meckley et al. (2014a, b). Prior to surgery, sea lamprey were anesthetized by immersion in 0.2 mL·L−1 clove oil solution. The anesthetic solution was produced through dilution of 2 mL of clove oil (minimum 84%-88% eugenol, Lot No. HB9387, Hilltech Canada Inc. Vankleak Hill. Ontario, Canada) into 18 mL of 70% ethanol and vigorously mixed into 10 L of Lake Huron water. Sea lamprey were removed from the bath upon reaching stage four of anesthesia, determined by individuals that did not respond to handling but retained gill movement (mean time to stage ± 1 SE, 559 ± 14.15 s, maximum 846 s). The surgery was performed in a PVC pipe with continuous water flow that allowed gill irrigation to be maintained by completely submerging the head and gills. We inserted the transmitter into the peritoneal cavity through a 20 mm incision approximately 10 mm off the ventral midline that ended in line with the anterior insertion of the first dorsal fin. The incision was closed with three independent interrupted surgeon knots (3-0 Ethicon sterile monocryl monofilament) and each knot was sealed with veterinary adhesive (Vetbond, n-butyl cyanoacrylate adhesive). The surgical procedure took an average of 283 ± 4 s to complete. Each subject was monitored in a postoperative holding tank until we observed the animal regain equilibrium and begin natural swimming movements (recovery time, mean ± 1 SE, 360 ± 28 s, maximum 1240 s). Transmitter-implanted lamprey were held for 72 h prior to release to ensure metabolism of stress compounds (Close et al. 2003). Data quality assessment and position filtering The position accuracy of the VPS array was tested by comparing the VPS position estimates to GPS measured positions (Trimble Geo XH, post processed) of two transmitters at fixed locations (Fixed test 1: June 17, 2014 – June 26, 2014; Fixed test 2: June 17, 2014 – July 01, 2014) and two transmitters pulled through the array (Drag 1:June 13, 2014; Drag 2: June 73 17,2014) (V9P-2H transmitter). To avoid the effects of positioning error on path sinuosity, a data quality objective was set to attain 95 % confidence that reversals did not erroneously occur in the data. VEMCO positioning systems provide a position precision estimate for each position, horizontal positioning error (HPE), if evaluated the HPE can be used to remove erroneous positions (Smith 2013, Meckley et al. 2014a, Roy et al. 2014). The step length was 22 m in calm wave conditions, which equated to an objective of 95% confidence in an error less than 11 m. We estimated the twice the distance root mean square error (2DRMS) equations for each fixed tag and calculated maximum HPE values that would allow 95 % confidence in 11 m of error. We selected an HPE filter of 15.17 (Appendix 1). To classify sea lamprey positions as active or stopped, a first passage-time classification method was performed with the program R. The first passage-time tool classified a position as moving if it left a radius of 10 meters in 250 s, had a minimum displacement of 15 m in a 3 position moving average, and at least 3 consecutive moving observations (function “fpt”, package adehabitatLT) (Calenge 2006, Gurarie et al. 2015). The assignment accuracy was verified for each sea lamprey through visual inspection and was robust to imprecision in acoustic positioning largely because we did not observe position error of greater than 10 m for three consecutive positions. Sea lamprey were assigned to one of 5 categories based on post-release behavior: (A) those that stopped in the array and exited the array on the first night (n=22); (B) those that stopped at the edge of the array and began on the first night but few active points were captured because they were near the edge of the detection range (n=5); (C) those that stopped in the array and exited on the second night or later (n=10); (D) those that immediately abandoned the array on the bottom after release (n=8); and (E) those that immediately abandoned the array after 74 release but moved throughout the water column (n=22). Only those in the predefined group A, that stopped before leaving the array on the first night were evaluated for orientation behavior and vertical movement patterns although we compared groups A and E in terms of their use of the water column when moving vertically at day and night as there was an apparent daytime avoidance of movement in the upper area of the water column. Sea lamprey are nocturnal and the internal state of the subjects that exited the array during the day immediately after release cannot be reliably interpreted as searching for the coast or searching for cover (Almeida et al. 2002, Vrieze et al. 2011, Meckley et al. 2014b). Only considering fish that followed a progression of settlement and inactivity during the day and activity at night protects our findings from adverse tagging effects commonly seen immediately after release (Frank et al. 2009). Statistical analyses (1) Do Sea lamprey perform y-axis orientation? To determine if sea lamprey approached the nearest southern coastline, we analyzed detection of tags at receivers on the coast within 72 hours of exiting the array. This timeframe ensured that sea lamprey could not have reached a northern coast prior to reaching the nearest southern coast. Lamprey observed on the southern coast, were classified as successfully orienting; the null expectation was that absent an orientation mechanism the probability of success would be 50%. (2) Are there two phases of movement? To determine if a transition occurred between two phases of movement and whether trajectory parameters differed between the postulated phases, we evaluated the paths of 22 sea 75 lamprey from the commencement of movement from a settled state until exiting the region of VPS array coverage. To classify the two phases of: i) apparent initial undirected (sinuous movement), followed by ii) directed movement (extensive straight movement), we analyzed the sinuosity of the path in terms of the progression of the backward beeline distance to total backward path length at each position along the track to the vanishing point for each sea lamprey (Bovet and Benhamou 1988, Girard et al. 2004). A broken stick model was fit to each path and optimized across four parameters including initial slope (𝛽1), break point (𝜏), sigma (σ), and final slope (𝛽2), using maximum likelihood estimation with function “optim” in program R (R Development Core Team 2015). In the model backward path length (𝑥) was the explanatory variable for estimating the response variable backward beeline distance (𝑦). The y intercept (𝛽0 ) was fixed at 0 due to the nature of the parameter always returning to (0, 0). Confidence intervals were estimated from the Hessian of the log-likelihood. The model used: 𝑦 = 𝑦̂ + ε; ε ~ Normal(0, σ2 ) 𝑦̂ = 𝛽1 𝑥1 + 𝛽2 𝑥2 + ε 𝑥1 = { 𝛽(𝑥), 𝜏, 0, 𝑥2 = { (𝑥 − 𝜏), 𝑥≤𝜏 𝑥>𝜏 𝑥≤𝜏 𝑥>𝜏 When 𝑥 ≤ 𝜏 the model is linear with slope 𝛽1 and when 𝑥 > 𝜏 the model linear with slope 𝛽2 and intercept constrained by continuity to the first part of the stick. In the special case of 𝛽1 = 𝛽2 the model is a simple linear relationship between x and y (White et al. 2008). The initial break point parameter provided was estimated through visual inspection of each of the 22 tracks. Individual pairwise t-tests were used to determine if trajectory parameters that are not directly associated with track sinuosity (e.g. ground speed, turning bias, average depth, and variation in 76 depth) varied between phase one (before break point) and phase two (after break point). We predict that if a distinct exploratory phase occurs, the initial behavior will be accompanied by slower more sinuous movement followed by straighter faster movement. (3) Is there evidence for orientation towards a consistent cardinal direction? We tested if there was a significant sample mean direction with a Raleigh test based on an unspecified mean (function “raleigh.test”, R package circular) or a significant non-uniform distribution via a Watson’s test (function “watson.test”, R package circular) for the portion of tracks between the transition point and departure point from the array (Agostinelli and Lund 2013). These tests are similar in nature and answer the same question by either using a z-test statistic (Raleigh test) or goodness of fit to a circular uniform distribution (Watson’s test). If sea lamprey orient to a consistent broadly available feature such as visual recognition of the closest coast, we predicted sea lamprey would have a significantly clustered mean direction and a nonuniform distribution. If sea lamprey were not orienting, oriented to multiple cues at different times, or oriented to a locally varying feature such as bathymetry in our array, we predicted sea lamprey would not have a significant mean heading or a uniform distribution in any compass direction. (4) Do sea lamprey orient directly to a coastal feature or the bathymetric gradient? To evaluate if sea lamprey orient to a feature of the closest coast or bathymetry, we evaluated the heading of each individual during phase 2, defined as the portion of tracks between the transition point and departure point from the array. We tested if sea lamprey moved towards the closest coast (coast test) or towards shallower water (bathymetry test). For the bathymetry 77 test, the bearing of the local bathymetric slope was first calculated from the bearing of the deepest to shallowest depth point on a 1000 point ring of positions in a 100 m radius around the final point to capture the general slope for each sea lamprey leading to the exit. After all bathymetric headings were verified visually on a map, the heading was subtracted from the phase two headings for each individual so that a bathymetry test heading of near 0 o (or 360 o) indicated that the sea lamprey was moving in the ideal bathymetric direction and 180 o represented moving the opposite direction. If sea lamprey oriented in the direction of the local bathymetric slope towards shallower water we predicted the Rayleigh test of the turn angle between fish heading and shallower slope heading with a specified mean turn angle of 0 would be significant and a Raleigh test of the fish heading with a specified mean of 172o for the direction to shallower water would be insignificant. (5) Did the river the sea lamprey was trapped in influence orientation? To determine if the lamprey’s capture river influenced orientation (experience), the distribution of Phase 2 headings between river sources was compared using a Watson-Wheeler test (“Watson.wheeler.test”, R package circular) and the differences in the distance traveled during Phase 1 were considered (two sample t-test) (Agostinelli and Lund 2013). We compared the sea lamprey trapped in the Manistique River and Cheboygan River. If the river they were trapped in had an effect, we would expect sea lamprey migrants from the Cheboygan River to reach the nearest coast after a shorter orientation process, and have a different distribution of directed headings. Although the Watson Wheeler test only tests for a difference between groups, our expectation was that the phase 2 heading would be more southerly towards the local coast, 78 than the Manistique source subjects. If a sea lamprey relied on local information to orient, we predicted no effect of river source on orientation. Characterizing the process: Do lamprey show consistent vertical excursions? To characterize the vertical movements of sea lamprey through the water column beyond average depth and standard deviation of depth, vertical occupancy was characterized for each step (two consecutive points) into one of three classes of vertical movement (ascent, descent, or horizontal phases) based on pitch (descent < -5o, horizontal ≥ -5o and ≤ 5o, ascent > 5o). Because sea lamprey move from surface to bottom during each phase, we compared vertical movement with the time spent moving horizontally within each sinuous phase, the depth of sea lamprey when moving horizontally in each phase, and the number of casts, defined as the number of times sea lamprey ascended and descended at least 10 m in succession. These variables were used to assess whether the vertical movements represented constant and consistent oscillations or primarily horizontal movements on the surface or bottom with intermittent movements to either the surface or bottom. Individual pairwise t-tests for each sea lamprey were checked for differences in depth between Phase 1 and 2 during horizontal movements (horizontal ≥ -5o and ≤ 5o). To test for differences in elapsed time spent moving horizontally in the water column during phases 1 and 2 and if a difference existed in the rate of vertical displacement or pitch during ascent and descent, we used a logistic regression analysis. A single response variable including time moving horizontally(s), pitch angle (o), or vertical displacement rate (m/s)), respectively. Each was fit using a generalized linear mixed effects modeling framework. For testing for differences in time spent moving horizontally, the movement phase was a fixed effect, individual sea lamprey 79 (identification number) was a random effect and we weighted the model by the number of observations. For testing displacement rates and pitch during ascent and descent, ascent and descent were fixed effects, while individual sea lamprey was a random effect. The explanatory variables, pitch and vertical movement rate, were considered positive values, regardless of whether the sea lamprey was ascending or descending in the water column. Finally the shallowest depth reached for sea lamprey that showed regular vertical movements during the day versus at night were tested with a simple two way t-test. This was the only test where data for sea lamprey that did not stop immediately after release were used. All analyses were performed in R (R Development Core Team 2015), including the lme4 package (Bates et al. 2014) for fitting the mixed effects model. RESULTS (1) Do Sea lamprey perform y-axis orientation? Evidence suggested that sea lamprey perform y-axis orientation, although only 22 of 67 released sea lamprey stopped in the array after release and left on the first night (Group A). An additional five individuals stopped at the edge of the array where only intermittent positions were attained and nocturnal activity was not recorded (Group B). Prior to stopping at the edge of the array, two fish made vertical excursions of greater than 10 m prior to stopping, and one fish moved laterally on the bottom; the remaining individual did not have depth data. Of 10 fish that did not exit the array on the first night, six were from the eight total Manistee River fish (Group C). Sea lamprey that did not stop included eight (Group D) individuals that exited the array on the bottom and 22 that showed vertical movements of at least 10 m off of the bottom (Group E). Of the 27 sea lamprey (Group A and Group B), 21 (78%) arrived at the nearest coastline within 72 hours and most that were detected at the nearest two receivers were detected on the first night 80 (12 of 15) (Table 3.1). If considering all tagged lamprey, 43 of the 67 (61%) sea lamprey were detected on the nearest coast (Table 3.1). (2) Are there two phases of movement? We observed a post-settlement orientation pattern consistent with initial sinuous movements (Phase 1) followed by straight movements (Phase 2) for 21 of the 22 sea lamprey (Figure 3.3, Figure 3.2). The remaining individual immediately departed in an apparent straight line without an exploratory phase (T26, Figure 3.2). Phase 1 persisted for an average track length of 808 m ± 178 m (mean ± 1SE) or an interquartile range of 409-837 m (1st IQR-3rd IQR). Phase 2 lasted 975 m ± 110 m with an Interquartile range of 625-1311 m. Several aspects of the sea lamprey movement varied between the classified phases, including persistence in turning in one direction, amount of turning, mean depth, variation in depth, and ground speed (Table 3.2, Table 3.3). During Phase 1, 10 of the 21total individuals turned with a left-bias while four individuals were right-biased and the magnitude of persistence in turning of biased individuals was greater before than after the transition (Table 3.2, Figure 3.4). Sea lamprey turned less during Phase 2 though this observation could be confounded by the fact that the break point was defined in terms of the path sinuosity (Table 3.2). Mean depth was shallower after the break point for 14 of 16 individuals and the standard deviation in depth was significantly greater for five of 16 individuals vs. significantly reduced for two of 16 individuals, who appeared to swim near the surface without surface to bottom casting in Phase 2. Lastly, ground speed was significantly faster for 18 of 21 individuals during phase two when compared to phase one (Table 3.4). 81 (3) Is there evidence for orientation towards a consistent cardinal direction? During Phase 2, sea lamprey assumed an average absolute heading of 239o ± 96o (mean ± 2STD, 95 % CI: 201-276 deg.; Figure 3.5). Sea lamprey did not show an absolute orientation in any cardinal direction (r = 0.24, p = 0.26; Rayleigh Test, unspecified mean) and their circular distribution was not significantly different from random (test statistic = 0.091, critical value at a significance of 0.05 = 0.187; Watson’s Test) (Figure 3.6). (4) Do sea lamprey orient directly to a coastal feature or the bathymetric gradient? No evidence indicated that sea lamprey oriented towards the nearest coast (r = 0.11, p = 0.22; Rayleigh Test, specified mean: 173 deg.). By contrast, Phase 2 headings (mean: 353 deg., 95% CI: 307-49 deg., were not randomly distributed with respect to bathymetry (Watson’s Test: test statistic= 2.244, critical value= 0.187, reject Null) and were significantly oriented to local bathymetry in the region where they exited the array (r = 0.32, p = 0.017; Rayleigh Test, specified mean: 0o). (5) Did the river from which the sea lamprey was trapped influence orientation? The Watson-Wheeler test of homogeneity of angles found no significant difference between sea lamprey trapped in different places, although the strength of this conclusion is limited because of the small sample size (N<10) for each group (W = 0.76, p = 0.69). No significant difference occurred between the distances traveled prior to the break point for Cheboygan River fish (mean ± 1SE, 972 m ± 207 m) or Manistique River fish (mean ± 1SE, 976 m ± 244 m) (p=0.99) . 82 Characterizing the process: Do lamprey show consistent vertical excursions? Sea lamprey moved vertically throughout the entire water column (e.g., Figure 3.7). On average, they made seven ascents or descents of at least 10 meters (6.7 ± 3.2 dives per h Phase 1, 7.3 ± 4.5 dives per h Phase 2). Sea lamprey maintained a straight course (no turns >15o) for > 500 m during the day, at night, and while moving at the surface, moving on the bottom, and while ascending and descending through the water column. Sea lamprey ascended at a pitch of 15.0 ± 7.8 and vertical speed of 5.1 ± 2.8 m·min-1 during Phase 1 and a pitch of 13.0 ± 6.8 o and vertical speed of 5.6 ± 2.5 m·min-1 during Phase 2. Sea lamprey descended at a pitch of 14.9.0 ± 8.0 o and vertical speed of 7.1 ± 7.9 m·min-1 during phase one and a pitch of 14.9 ± 7.7 o and vertical speed of 8.3 ± 5.2 m·min-1 during phase two. During Phase 1, sea lamprey primarily moved on the bottom and made occasional vertical excursions to the surface (Figure 3.8). During Phase 2, sea lamprey made more vertical excursions and spent more time at the surface than the bottom (Table 3.4). Mixed effects logistic regression modeling revealed that sea lamprey spent significantly less time moving horizontally during Phase 2 than Phase 1 (estimate: - 0.42, p < 0.01), and more time moving vertically through the water column. Mixed effects logistic regression revealed that no difference occurred in the pitch of sea lamprey during ascent or descent (p = 0.08), but a difference did exist in the vertical rate of ascent and descent (p < 0.001). A two way t-test revealed that vertical excursions during the day did not extend as close to the surface as at night, rarely entering the upper 10 meters of the water column during the day (p: <0.01; mean minimum depth ± SD; Day: 10.4 m ± 5.4m, Night: 2.6 m ± 3.6 m). DISCUSSION The offshore movement of sea lamprey was consistent with y-axis orientation to the local bathymetric gradient towards shallower water; we will refer to this class of search as 83 bathokinesis. We addressed this complex hypothesis through a bundle of predictions. (1) The majority of sea lamprey (78%) arrived at the nearest coast following release. (2) Sea lamprey (21/22) displayed a two-phase pattern that varied in sinuosity consistent with orientation (Phase 1) followed by directed search (Phase 2). During Phase 1, sea lamprey moved slowly on the bottom and turned, often in a consistent direction. These tracks included vertical ascents to the surface and returns to the bottom, building from partial to full excursions through the water column. During Phase 2, sea lamprey moved more quickly and steered a straight course while making occasional heading corrections and small looping turns (pirouettes), and moved primarily at the surface, with occasional descents to the bottom that returned to the surface. (3) Sea lamprey did not orient in a consistent cardinal direction, signifying a variable search strategy or use of a varying local feature. (4) Sea lamprey did move perpendicular to the local bathymetric contours to shallower water. (5) The orientation process and resulting heading was not different for sea lamprey trapped in different rivers. In contrast to a natal homing migration, the nonhoming sea lamprey migration joins a growing number of organisms that orient to a region of resource availability with a simple set movement rules. These findings revealed that sea lamprey oriented consistently with the bathymetric slope, though the findings did not indicate an orientation mechanism or how sea lamprey maintained a straight course. We postulated that these sea lamprey: i) used the absolute hydrostatic pressure sensed at successive sites along the lake bottom to pick a heading towards shallower water during Phase 1, and ii) used declining hydrostatic pressure at the lake bottom to steer towards shallower water during Phase 2. Because sea lamprey moved straight during phase two regardless of their position in the water column, it would be reasonable to infer that sea lamprey rely on additional directional features to maintain course, rather than relying on a 84 gradient based feature of the environment. We have identified four possible mechanisms for how sea lamprey follow the bathymetric gradient based on the characteristics of the physical water column and the nature of sea lamprey movement. Sea lamprey could orient to a feature of bathymetry when moving at the bottom such as absolute hydrostatic pressure, a physical aspect of the bottom substrate or bottom pitch, or by moving vertically through the water column and using a vertically collected feature like dynamic hydrostatic pressure or the physical extent (e.g., time or distance) traversed through the water column (Taylor et al. 2010). An internal clock method keeping track of the time or distance between the surface and bottom was not supported as sea lamprey did not maintain pitch and vertical movement speed. Vertical speed was significantly faster when descending and pitch was nearly significant (p=0.08) shallower during ascent. Bottom substrate did not contain features like sand waves found in shallower water that would inherently indicate the direction to the coast (Figure 3.9). Fishes have been shown to orient to hydrostatic pressure (Cain 1995, Holbrook and Burt de Perera 2009, Holbrook and Burt de Perera 2013), and can discern differences in absolute or differential hydrostatic pressure using the lateral line, inner ear (labyrinth), and swim bladder (Fraser 2002, Fraser et al. 2008, Bleckmann and Zelick 2009). Sea lamprey lack a swim bladder and have a primitive lateral line and inner ear that lacks a lagena (otolith endorgan); however, sea lamprey possess semicircular canals that are structurally similar to the otolith of other pressure sensitive fishes (Hammond and Whitfield 2006, Fraser et al. 2008, Khorevin 2008), which can detect changes on the order of 0.5-2.0 kPa (5-20 cm) (Blaxter 1980). We postulate that sea lamprey use absolute hydrostatic pressure at the maximum depth to orient towards shallower water through turning and correcting towards shallow water (bathokinesis), and short looping pirouettes during the directed phase to reaffirm their selected heading. The movement 85 pattern was consistent with kinesis rather than taxis, as initial headings were undirected (appearing random) and turns were made until movement occurred towards shallower water (Miller et al. 2009). The purpose of excursions to the surface is unclear and introduces two challenges to using hydrostatic pressure to orient; first, sea lamprey must be capable of recalling previous dives and second they must be able to maintain a course when moving off of the bottom. Some fishes are capable of remembering information in three-dimensional space to locate targets (Holbrook and Burt de Perera 2013) and in open-water environments lacking horizontal edges (walls), the vertical component of the water column may be more informative to navigation (Holbrook and Burt de Perera 2009, Holbrook and Burt de Perera 2013). Recalling the maximum hydrostatic pressure would allow for orientation but maintenance of a straight path requires a second directional feature. Two common navigational mechanisms include reliance on geomagnetic fields or current direction to maintain course. Whether sea lamprey are capable of geomagnetic navigation is unknown, though this form of navigation is commonly used by other fishes that make large migrations and should be tested in this basal vertebrate. Water currents in this environment are highly variable vertically and horizontally and usually weak (< 2 cm/s) (Beletsky et al. 1999), making for a potentially poor directional cue in the lake. On the other hand, current is strongest at the surface (Beletsky and Schwab 2001), a potential reason for moving to the surface. Combining this with knowledge that sea lamprey do not vertically cast in deep rivers (16 m depth, Holbrook et al. 2015), but do vertically cast in the presence of river water (i.e., Meckley et al. 2014b), we propose that vertical movement is designed to encounter stronger currents, or navigate absent strong currents. 86 Animals could make vertical excursions to improve migration performance, to use navigation cues located at the surface and bottom, to search for a missing cue stratified in the water column, or to avoid sensory habituation to a cue (Westerberg 1982, Klimley et al. 2002). We had no evidence for an improvement in migratory performance as the movement increases the total path length and sea lamprey do not spend more time falling than rising as would likely be required to receive an energetic advantage (Carey et al. 1990, Klimley et al. 2002). The occurrence of vertical excursions in a shallow fully mixed water column (Vrieze et al. 2011, Meckley et al. 2014b), suggests against the presence of a thermoregulatory advantage (Katz 2002). Sensory habituation can occur if a receptor is constantly exposed to a stimulus and becomes fatigued resulting in a loss of sensitivity (Ferrari et al. 2010). In theory, vertical excursions in and out of a cue could prevent sensory habituation. Sea lamprey could encounter information that is stratified at the surface or when present would reside at the surface (missing cues). River water can be thermally pinned at the lake surface or bottom of the water column (Masse and Murthy 1990, Churchill et al. 2003), or mixed throughout the water column, which led Vrieze et al. 2011) to conclude that sea lamprey vertically cast to encounter olfactory evidence of a river. This is a plausible explanation, though it is likely that additional navigational benefits are gained by moving vertically (Meckley et al. 2014b), but from our observations we cannot report a dominant purpose. The sea lamprey migration in the Great Lakes appears to rely on a simple set of rules. Sea lamprey follow the local bathymetry to return to the coast, potentially relying on absolute hydrostatic pressure. Once arriving at the coast, they move parallel to the coast until encountering a river plume, and then transition to deflecting along the local coast around the river mouth to ensure encounter with the river outflow, as they assess or enter the river (Vrieze et 87 al. 2011, Meckley et al. 2014b). The leading hypothesis is that once encountering a river plume individuals either enter, stage, or search for a different river based on the presence of larval odor (quality) and river temperature (timing)(Binder et al. 2010, Clemens et al. 2010, Meckley et al. 2012). The odor of river water and larvae from previous generations is attractive to sea lamprey and is surmised to inform sea lamprey of the presence of a river mouth and the quality of the spawning and rearing habitat upstream (Moore and Schleen 1980, Bjerselius et al. 2000, Vrieze and Sorensen 2001). Simple orientation rules based on features of the geophysical environment that can reliably lead to a target rich region appear to be a common attribute of non-homing animal search, rather than reliance on direct information associated to a target as is found in homing migrations. Future Directions As commonly occurs in novel studies of animal migration, we identified two hypotheses that should be tested further and additional features of the sea lamprey migration worth consideration. First, we postulate that sea lamprey can use gradients of hydrostatic pressure at the bottom of the water column to determine the bathymetric gradient which provides them the direction towards shallower water. Second, we postulate that sea lamprey move to the surface to gather other navigational information and/or to encounter additional cues (e.g. chemical, current) that could be structurally stratified at some point throughout the water column, while also preventing habitation to sensory cues in the environment. Unexpectedly a larger proportion of sea lamprey in this study did not stop immediately after release and still moved vertically through the water column (33 %), a finding conflicting with two years of study that found in shallow water (<10 m) in both years only 17 % of sea 88 lamprey didn’t stop and the movement was nearly exclusively on the bottom (unpublished data). Though in this study, when sea lamprey did move after release they did not move into the upper 10 meters of the water column during the day but moved all the way to the surface at night. It remains unclear if sea lamprey are more active during daylight hours in deeper water (>10 m). During the open water migration if sea lamprey avoid the upper water column, the strike range of the sea lamprey’s most documented predator (avian predators, Close et al. 2002) could be avoided. It is also possible that this observation was just an artifact of being released during the day coupled with an inherent aversion to strong light and less inclination to stop immediately in deeper darker water. Three other observations of uncertain relevance occurred including the persistent turning that created loops in a consistent direction during phase one, the observation that few fish went east after reaching the coast, and the failure of fish from more southerly rivers to exit the array. The handedness of turning appeared to be dominated by counterclockwise turners, which could be an artifact of our small data set or exist due to an ecological advantage to handedness. Handedness in turning during search has been suggested to improve energetic efficiency (e.g., Kells and Goulson 2001), making the behavior worth additional consideration. Secondly, fish appeared less likely to go east after reaching the coast (Table 3.1). This directionality could be the result of moving opposite the dominant lake current in this region of the lake (Beletsky et al. 1999) or some other bias. The final interesting feature of the study came from the failure of sea lamprey from more southerly rivers to exit the array on the first night. These individuals did leave the array after a period of several days. These individuals were caught farther upstream prior to tagging, and later in the year than the other individuals, and had visibly more developed 89 gonad. It appeared that these individuals required some type of reset period before they began to migrate out of the array. Bathokinesis should be examined in ocean-run sea lamprey. Both Pacific lamprey and sea lamprey have been are neither panmictic nor philopatric (Waldman et al. 2008, Spice et al. 2012, Hess et al. 2013), suggesting that some aspect of their life cycle constrains them to a region but also disperses them. Additional evidence of restricted movement exists, as North American and European populations do not mix (Rodriguez-Munoz et al. 2004, Genner et al. 2012), and move broadly throughout the ocean when parasitizing other fishes, in some cases moving over 815 km from the nearest coast (Silva et al. 2014). Additional mechanisms may occur to prevent navigational mistakes that result in the mixing of populations within the Atlantic basin sea lamprey or some control over when a parasite will release from a host to prevent large displacements. If sea lamprey select hosts that migrate back to rivers, it is plausible that sea lamprey could hitch a ride back to rivers in some instances; a behavior termed phoresy, although we suspect this is not the dominant strategy. The only evidence available is consistent with bathokinesis by ocean-run sea lamprey, as parasitic sea lamprey feeding within a general basin were constrained by basin topography during their return migration (Lanca et al. 2014). ACKNOWLEDGEMENTS This research was funded through grants from the Great Lakes Fishery Commission including by way of the Great Lakes Restoration Initiative appropriations (GL-00E23010-3) and was accomplished through considerable assistance from the US Fish and Wildlife Service and the USGS Hammond Bay Biological Station. 90 APPENDIX 91 Table 3.1: Sea lamprey were primarily detected on the nearest coast within 72 hours of release. This included 78 % of individuals that were observed stopping in the array and moving on the first night (n=22) and 61% of all individuals released (n=67). Note that Time to reach the receiver mean and min/max are only listed for the 22 individuals. Only 4 of 67 sea lamprey’s first detections occurred outside of the nocturnal movement period, with 3 of 4 occurring on the first day of release with 2 at Forty Mile point (East), 1 at the Ocquoec river mouth (South) and 1 near Cheboygan river mouth (West). No. of Receiver/ Distance Time to reach Time to reach unique detection from array receiver hours receiver hours individuals Site (km) (mean ±SE) (min, max) (N=22, N=67) OCQ 4.3 (9, 17) (6.9 ± 5.0) (0.83 - 46.7) (12, 26) BM 6.2 (6, 11) (17.7 ± 10.7) (1.7 – 69.3) (9, 17) CHB 33.8 (2, 9) (51.8 ± 0.4) (51.4 - 52.2) (4, 15) SMR 51.6 (2, 5) (141.0 ± 86.3) (54.7, 227.3) (2 , 5) FMP 11.1 (0, 4) (1, 5) PRS 48.9 (0, 0) (1, 1) The number of sea lamprey detected on any receiver was 19 of 22 and 46 of 65). Ten of 23 and 19 of 65 sea lamprey entered the Ocquoec River. 1 of 22 and 2 of 65 entered the Black Mallard River. Four of 23 and seven of 67 sea lamprey were trapped in the Ocqueoc River and one of 67 was trapped in the Cheboygan River. No. of 1st Detections in 72 hours (N=22, N=67) 92 Table 3.2: Individual t-tests for each sea lamprey of whether sea lamprey exhibited a persistence in turning (theta; -pi to pi), different from 0, revealed that more sea lamprey had a bias of left (negative) or right (positive) turns during phase 1 (P1), but not during phase 2 (P2). Most persistent turns were counterclockwise. The magnitude is also greater for those that are significant in phase 1 than phase 2. A pairwise t-test of the concentration in turning (Rho) for each individual between phase 1 and phase 2 revealed that individuals tended to go straighter during phase 2. Bold values are significant. Theta ID P1 T02 -0.03 T04 -0.06 T05 -0.17 T07 -0.81 T08 0.07 T12 -0.58 T17 0.14 T22 -0.43 T25 0.50 T27 -0.26 T31 -0.47 T32 0.07 T36 -0.58 T37 0.02 T42 0.14 T44 -0.28 T47 0.68 T54 -0.12 T56 -0.13 T61 -0.01 T63 -0.20 Total ( ρ < 0.05): Total Possible: P1 ρ P2 P2 ρ 0.488 0.494 0.214 0.000 0.689 0.000 0.625 0.001 0.001 0.004 0.001 0.792 0.000 0.878 0.082 0.836 0.028 0.107 0.006 0.903 0.005 10 21 -0.08 0.00 0.01 -0.21 0.04 0.06 -0.10 -0.04 -0.07 -0.06 0.14 0.05 0.01 0.33 -0.14 0.03 -0.08 0.01 -0.18 0.06 -0.07 0.04 1.00 0.79 0.29 0.46 0.48 0.48 0.51 0.27 0.67 0.24 0.52 0.88 0.00 0.25 0.61 0.30 0.76 0.01 0.56 0.02 4 21 93 Concentration of turns (Rho) P1 P2 ρ 0.81 0.70 0.74 0.74 0.48 0.61 0.54 0.59 0.64 0.61 0.63 0.65 0.67 0.69 -0.28 0.62 0.72 0.69 0.75 0.65 0.83 0.86 0.86 0.65 0.84 0.77 0.75 0.88 0.81 0.75 0.73 0.69 0.75 0.85 0.78 0.79 0.84 0.81 0.72 0.73 0.86 0.61 0.01 0.18 0.47 <0.01 0.18 0.66 <0.01 0.01 0.18 0.36 0.77 0.19 0.12 0.34 0.21 0.14 0.10 0.37 0.82 <0.01 5 21 Table 3.3: Pairwise t-tests to determine if there was a difference in the mean depth, standard deviation in depth, mean depth when Mean Depth (m) ID T02 T04 T05 T07 T08 T12 T17 T22 T25 T27 T31 T32 T36 T37 T42 T44 T47 T54 T56 T61 T63 P1 P2 32.85 17.98 32.69 15.85 29.80 21.59 33.86 23.76 35.07 28.09 21.38 8.04 23.52 14.67 22.72 4.97 32.91 5.99 23.16 7.51 27.31 13.95 30.48 11.31 31.98 31.06 28.93 26.25 15.02 3.84 26.67 5.94 Total Significant (p < 0.05): Total Possible: ρ <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 0.09 <0.01 <0.01 <0.01 <0.01 <0.01 0.13 0.04 <0.01 <0.01 Standard Deviation Depth (m) P1 P2 ρ 4.86 9.38 0.00 6.60 9.89 0.08 8.79 11.45 0.38 0.15 8.30 0.00 0.92 6.84 0.00 10.54 6.69 0.14 8.42 10.96 0.62 12.08 3.62 0.00 3.90 3.85 0.95 11.88 6.93 0.03 6.51 11.37 0.13 3.44 9.39 0.00 1.79 5.61 0.00 2.90 3.11 0.87 9.06 2.14 0.06 9.94 7.45 0.28 Hz Mean Depth (m) P1 35.1 36.8 34.1 33.9 35.3 26.5 32.0 24.2 34.4 25.0 25.8 21.5 32.1 30.2 8.6 - P2 14.6 7.1 20.3 24.8 27.5 4.4 6.2 3.8 4.8 3.9 4.7 3.4 32.5 26.1 2.3 4.5 Ground Speed (m/s) ρ 0.05 <0.01 0.16 0.21 0.23 <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 0.12 0.35 0.28 - P1 0.38 0.38 0.38 0.33 0.29 0.33 0.59 0.30 0.17 0.35 0.34 0.19 0.36 0.40 0.42 0.19 0.56 0.37 0.37 0.46 0.27 P2 0.49 0.48 0.49 0.50 0.51 0.51 0.69 0.52 0.30 0.47 0.57 0.40 0.43 0.44 0.73 0.44 0.69 0.60 0.52 0.51 0.53 ρ <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 0.08 <0.01 <0.01 <0.01 <0.01 <0.01 <0.01 0.15 <0.01 <0.01 0.08 <0.01 <0.01 <0.01 <0.01 12 7 7 18 16 16 14 21 94 Table 3.3: (cont’d) maintaining vertical depth (Hz Mean Depth), or ground speed during phase 1(P1) and before the transition to phase 2 (P2). The tests support our observations that these two phases are different. Sea lamprey had greater mean depths and less variation during phase 1, and when moving at a particular depth in the water column, sea lamprey spent time on the bottom during phase 1 and closer to the surface during phase 2. The ground speed (GS) was also significantly lower during phase 1 than phase 2. The only individuals with less variation in depth during phase two were those individuals swimming mostly at the surface with occasional vertical excursions. Individuals T42-T56 did not have pressure sensitive tags. 95 Figure 3.1: Adult female sea lamprey were obtained from 5 rivers via barrier traps (Manistee River (2a), Cheboygan River (1), Manistique River (3), Betsie River (2b), Ocquoec River (“OCQ”), and parasites were caught attached to fish in the lake. A telemetry system (VPS, VEMCO) composed of 41 receivers was used to position sea lamprey (“array”). There were individual receivers that provided detection data along multiple locations on the coast and in front of as well as in the Black Mallard (BM), Ocquoec (OCQ) and, St. Mary’s Rivers (SM). Receiver strings were treated as individual detection points and were located near the Cheboygan River mouth (CHB), 40 mile point (FMP), and Presque Isle (PRS). 96 Figure 3.2: Each sea lamprey was released near the bottom in the middle (yellow star) of a 43 receiver array (orange dots) with 3 km2 coverage. A receiver in the center of the array at the release point and on the east side of the array were not recovered, resulting in an irregular design. The black arrow demonstrates the direction to the nearest coastline 172o (3 radians). 97 Figure 3.3: The track of sea lamprey T04 is depicted as it exits the array (a), and below is the relationship between backward path length and backward beeline distance of the track (b). An arrow is drawn from the breakpoint to the end of the phase two to demonstrate the direction headed during Phase 2 (a). In the lower graph the yellow line is a one to one relationship from 0. A broken stick model fit is shown with the red lines representing the standard error and the black line is the line fit (b). The break point separates phase one (orange) from phase two (blue). The 22 tracks of Group A are shown in Figure 3.10. 98 Figure 3.4: Dotted lines connect phase 1 (orange) and phase 2 (blue), data points for each sea lamprey subject, separated by a transition characterized by backward path length versus backward beeline distance sinuosity. Both the mean value is shown (dot) as well as the standard deviation in each axis (+). There is an increase in ground speed and a transition from a turning bias in one direction to no bias in turning direction (a), as turning persistence overlaps 0 during phase 2. Also apparent was a change in the mean depth from mostly movement on the bottom in phase one with brief vertical excursions to the surface followed by movement on the surface during phase two and brief vertical excursions to the bottom (b). 99 Figure 3.5: The overall heading of 22 sea lamprey during phase two is depicted as a black dot within one of 60 bins separating the circle and shows an average heading of 239o (gray arrow) with a 95% confidence interval of 201-276o (shaded area) (a). The heading of sea lamprey with respect to the ideal heading if following local bathymetry at the exit of the array is depicted by the gray arrow with most sea lamprey clustered near 0 (mean= -7, 95% CI: 307-49o), repsenting a movement towards shallower water, while a few individuals were observed moving opposite of the ideal bathyemtric controur. The phase two headings were significantly oriented to local bathymetry in the region where they exited the array (r = 0.32, p = 0.017; Rayleigh Test, specified mean: 0o). 100 Figure 3.6: Sea lamprey (n=22) settled in the array during the day and began moving at night (yellow), transitioning (green circle) between phase one (red arrow) and phase two (black arrow) before they exited the array. Sea lamprey that moved north (b), east (c), south (d), and west (e) are shown in individual frames with symbols indicating where each individual was first detected on the coast. The array is color coded by depth 2039(m). 101 Figure 3.7: The track of sea lamprey T04 is depicted as it exited the array (a), and below is the water column depth (b) and ground speed during the track (c). The graphs are color coded by 15 minute intervals. In inset “a” small circles represent receiver positions and the small colored dots represent fish positions with a line showing the path. In inset “b” the red line represents the total water column depth and the squares show the depth of the fish with respect to the total water column depth through time.. All 22 tracks are shown in Figure 3.10. 102 Figure 3.8: Overall most observations occurred near the bottom (All observations; left panel). During phase 1 most sea lamprey moved on the bottom and made occasional excursions to the surface that returned to the bottom (center panel). Phase 2 consisted of more surface movements with occasional excursions to the bottom, although a few individuals did move vertically and only moved on the bottom (right panel). 103 Figure 3.9: Near shore sand waves in shallow water of Lake Huron (a) versus offshore detritus and algae covered bottom without clear features associated to a common direction (b). 104 Figure 3.10: There are 2 pages dedicated to each subject and only the data is shown for the points when fish began to move following settlement, with the exception of transmitter 21 that was shown moving immediately after release. The two pages correspond to the pathways of 22 subjects. Page one, “Movement Summary”, depicts the path of each sea lamprey through the array when active (a) and is color coded by 15 minutes intervals corresponding to the lower graph (b,c). Figure 3.10: (cont’d) Time is in Eastern Standard Time (UTC-4) and shown in a 24 105 Figure 3.10: (cont’d) hour time scale. The lower graph depicts the vertical movements of sea lamprey for those individuals that had transmitters with pressure sensors corresponding to the left y axis (b) and ground speed corresponding to the right y axis for all individuals (c). The red line is the total water column depth at each fish position. Note only T21 is shown moving during the day as it was notable because it appears to orient during the day and then repeated the process 106 Figure 3.10: (cont’d) at night but the night time track was too short to consider in the movement classification. This suggests that this is a daily orientation process. Page 2, “Movement Classification”, the lower graph depicts the backward path length versus the backward beeline distance for each track. The point at (0, 0) represents the point that the subject left the array. The yellow line is a one to one 107 Figure 3.10: (cont’d) relationship from 0. A perfectly straight path from the exit would move along this yellow line. The broken stick model fit is shown with the black line and the red lines represent the standard error. The break point is the point we use to separate phase one from phase two. Phase one is highlighted in the in orange and phase two is depicted in blue. In the upper graph, the path of each individual is shown and an arrow is drawn from the breakpoint to the end of phase two 108 Figure 3.10: (cont’d) when the fish exits the array coverage to demonstrate the heading of phase two that was used to test against the direction to shallower water and the direction to the nearest coast using Raleigh tests. 109 Figure 3.10: (cont’d) 110 Figure 3.10: (cont’d) 111 Figure 3.10: (cont’d) 112 Figure 3.10: (cont’d) 113 Figure 3.10: (cont’d) 114 Figure 3.10: (cont’d) 115 Figure 3.10: (cont’d) 116 Figure 3.10: (cont’d) 117 Figure 3.10: (cont’d) 118 Figure 3.10: (cont’d) 119 Figure 3.10: (cont’d) 120 Figure 3.10: (cont’d) 121 Figure 3.10: (cont’d) 122 Figure 3.10: (cont’d) 123 Figure 3.10: (cont’d) 124 Figure 3.10: (cont’d) 125 Figure 3.10: (cont’d) 126 Figure 3.10: (cont’d) 127 Figure 3.10: (cont’d) 128 Figure 3.10: (cont’d) 129 Figure 3.10: (cont’d) 130 Figure 3.10: (cont’d) 131 Figure 3.10: (cont’d) 132 Figure 3.10: (cont’d) 133 Figure 3.10: (cont’d) 134 Figure 3.10: (cont’d) 135 Figure 3.10: (cont’d) 136 Figure 3.10: (cont’d) 137 Figure 3.10: (cont’d) 138 Figure 3.10: (cont’d) 139 Figure 3.10: (cont’d) 140 Figure 3.10: (cont’d) 141 Figure 3.10: (cont’d) 142 Figure 3.10: (cont’d) 143 Figure 3.10: (cont’d) 144 Figure 3.10: (cont’d) 145 Figure 3.10: (cont’d) 146 Figure 3.10: (cont’d) 147 Figure 3.10: (cont’d) 148 REFERENCES 149 REFERENCES Able, K. P. (1991). "Common Themes and Variations in Animal Orientation Systems." American Zoologist 31(1): 157-167. Agostinelli, C. and U. Lund (2013). R package 'circular':circular statistics. R package version 0.4-7. Åkesson, S. and A. Hedenström (2007). "How Migrants Get There: Migratory Performance and Orientation." BioScience 57(2): 123-133. Akima, H. (1978). "A Method of Bivariate Interpolation and Smooth Surface Fitting for Irregularly Distributed Data Points." ACM Trans. Math. Softw. 4(2): 148-159. Almeida, P. R., B. R. Quintella and N. M. Dias (2002). "Movement of radio-tagged anadromous sea lamprey during the spawning migration in the River Mondego (Portugal)." Hydrobiologia 483(1-3): 1-8. Bates, D., M. Maechler, B. M. Bolker and S. Walker (2014). lme4: Linear mixed-effects models using Eigen and S4. R package version 1.1-7. Beletsky, D., J. H. Saylor and D. J. Schwab (1999). "Mean Circulation in the Great Lakes." Journal of Great Lakes Research 25(1): 78-93. Beletsky, D. and D. J. Schwab (2001). "Modeling circulation and thermal structure in Lake Michigan: Annual cycle and interannual variability." Journal of Geophysical Research: Oceans 106(C9): 19745-19771. Benhamou, S. (2006). "Detecting an orientation component in animal paths when preferred direction is individual-dependent." Ecology 87(2): 518-528. Bergstedt, R. A. and J. G. Seelye (1995). "Evidence for Lack of Homing by Sea Lampreys." Transactions of the American Fisheries Society 124(2): 235-239. Biesinger, Z., B. Bolker, D. Marcinek, T. Grothues, J. Dobarro and W. Lindberg (2013). "Testing an autonomous acoustic telemetry positioning system for fine-scale space use in marine animals." J Exp Mar Biol Ecol 448: 46 - 56. 150 Binder, T. R. and D. G. McDonald (2007). "Is there a role for vision in the behaviour of sea lampreys (Petromyzon marinus) during their upstream spawning migration?" Canadian Journal of Fisheries and Aquatic Sciences 64(10): 1403-1412. Binder, T. R., R. L. McLaughlin and D. G. McDonald (2010). "Relative Importance of Water Temperature, Water Level, and Lunar Cycle to Migratory Activity in Spawning-Phase Sea Lampreys in Lake Ontario." Transactions of the American Fisheries Society 139(3): 700-712. Bjerselius, R., W. Li, J. H. Teeter, J. G. Seelye, P. B. Johnson, P. J. Maniak, G. C. Grant, C. N. Polkinghorne and P. W. Sorenson (2000). "Direct behavioral evidence that unique bile acids released by larval sea lamprey (Petromyzon marinus) function as a migratory pheromone." Canadian Journal of Fisheries and Aquatic Sciences 57: 557-569. Blaxter, J. H. S. (1980). The Effect of Hydrostatic Pressure on Fishes. Environmental Physiology of Fishes. M. A. Ali, Springer US. 35: 369-386. Bleckmann, H. and R. Zelick (2009). "Lateral line system of fish." Integrative Zoology 4(1): 1325. Bovet, P. and S. Benhamou (1988). "Spatial-analysis of animal movements using a correlated random-walk model " Journal of Theoretical Biology 131(4): 419-433. Cain, P. (1995). "Navigation in Familiar Environments by the Weakly Electric Elephantnose Fish, Gnathonemus petersii L. (Mormyriformes, Teleostei)." Ethology 99(4): 332-349. Calenge, C. (2006). "The package "adehabitat" for the R software: A tool for the analysis of space and habitat use by animals." Ecological Modelling 197(3-4): 516-519. Carey, F. G., J. V. Scharold and A. J. Kalmijn (1990). "Movements of blue sharks (Prionace glauca) in depth and course." Marine Biology 106(3): 329-342. Churchill, J. H., E. A. Ralph, A. M. Cates, J. W. Budd and N. R. Urban (2003). "Observations of a negatively buoyant river plume in a large lake." Limnology and Oceanography 48(2): 884-894. Clemens, B. J., T. R. Binder, M. F. Docker, M. L. Moser and S. A. Sower (2010). "Similarities, Differences, and Unknowns in Biology and Management of Three Parasitic Lampreys of North America." Fisheries 35(12): 580-594. 151 Close, D. A., M. S. Fitzpatrick and H. W. Li (2002). "The ecological and cultural importance of a species at risk of extinction, Pacific lamprey." Fisheries 27(7): 19-25. Cooke, S., J. Midwood, J. Thiem, P. Klimley, M. Lucas, E. Thorstad, J. Eiler, C. Holbrook and B. Ebner (2013). "Tracking animals in freshwater with electronic tags: past, present and future." Anim Biotelem 1: 5. Craig, P. C. (1973). "Orientation of the sand-beach amphipod, ." Animal Behaviour 21(4): 699706. Dittman, A. and T. Quinn (1996). "Homing in Pacific salmon: mechanisms and ecological basis." The Journal of Experimental Biology 199(1): 83-91. Espinoza, M., T. Farrugia, D. Webber, F. Smith and C. Lowe (2011). "Testing a new acoustic telemetry technique to quantify long-term, fine-scale movements of aquatic animals." Fish Res 108(2): 364 - 371. Ferguson, D. E. (1971). "Sensory basis of orientation in amphibians." Annals of the New York Academy of Sciences 188(NDEC): 30-&. Ferguson, D. E., H. F. Landreth and M. R. Turnipseed (1965). "Astronomical Orientation of the Southern Cricket Frog, Acris gryllus." Copeia 1: 58-66. Ferrari, M. C. O., C. K. Elvidge, C. D. Jackson, D. P. Chivers and G. E. Brown (2010). "The responses of prey fish to temporal variation in predation risk: sensory habituation or risk assessment?" Behavioral Ecology. Frank, H. J., M. E. Mather, J. M. Smith, R. M. Muth, J. T. Finn and S. D. McCormick (2009). "What is "fallback"?: metrics needed to assess telemetry tag effects on anadromous fish behavior." Hydrobiologia 635(1): 237-249. Fraser, E. J., and Stacey, N. E. (2002). "Isolation increases milt production in goldifish." Journal of Experimental Biology 293: 511-524. Fraser, P. J., S. F. Cruickshank, R. L. Shelmerdine and L. E. Smith (2008). "Hydrostatic Pressure Receptors and Depth Usage in Crustacea and Fish." Navigation 55(2): 159-165. 152 Gebhardt, A., T. Petzoldt and M. Maechler (2013). akima: Interpolation of irregularly spaced data. R Package Version 0.5-11. Genner, M. J., R. Hillman, M. McHugh, S. J. Hawkins and M. C. Lucas (2012). "Contrasting demographic histories of European and North American sea lamprey (Petromyzon marinus) populations inferred from mitochondrial DNA sequence variation." Marine and Freshwater Research 63(9): 827-833. Girard, C., S. Benhamou and L. Dagorn (2004). "FAD: Fish Aggregating Device or Fish Attracting Device? A new analysis of yellowfin tuna movements around floating objects." Animal Behaviour 67: 319-326. Gurarie, E., C. Bracis, M. Delgado, T. D. Meckley, I. Kojola and C. M. Wagner (2015). "A comparison of methods and practical guide to the behavioral analysis of animal movements." Journal of Animal Ecology Special Issue. Hammond, K. L. and T. T. Whitfield (2006). "The developing lamprey ear closely resembles the zebrafish otic vesicle: otx1 expression can account for all major patterning differences." Development 133(7): 1347-1357. Hess, J. E., N. R. Campbell, D. A. Close, M. F. Docker and S. R. Narum (2013). "Population genomics of Pacific lamprey: adaptive variation in a highly dispersive species." Molecular Ecology 22(11): 2898-2916. Holbrook, C. M., R. Bergstedt, N. S. Adams, T. W. Hatton and R. L. McLaughlin (2015). "FineScale Pathways Used By Adult Sea Lampreys during Riverine Spawning Migrations." Transactions of the American Fisheries Society 144(3): 549-562. Holbrook, R. I. and T. Burt de Perera (2009). "Separate encoding of vertical and horizontal components of space during orientation in fish." Animal Behaviour 78(2): 241-245. Holbrook, R. I. and T. Burt de Perera (2013). "Three-dimensional spatial cognition: freely swimming fish accurately learn and remember metric information in a volume." Animal Behaviour 86(5): 1077-1083. Holland, R. A., I. Borissov and B. M. Siemers (2010). "A nocturnal mammal, the greater mouseeared bat, calibrates a magnetic compass by the sun." Proceedings of the National Academy of Sciences 107(15): 6941-6945. 153 Jonsen, I. D., R. A. Myers and J. M. Flemming (2003). "Meta-analysis of animal movement using state-space models." Ecology 84(11): 3055-3063. Katz, S. L. (2002). "Design of heterothermic muscle in fish." Journal of Experimental Biology 205(15): 2251-2266. Keefer, M. L., C. C. Caudill, C. A. Peery and M. L. Moser (2013). "Context-dependent diel behavior of upstream-migrating anadromous fishes." Environmental Biology of Fishes 96(6): 691-700. Kells, A. and D. Goulson (2001). "Evidence for Handedness in Bumblebees." Journal of Insect Behavior 14(1): 47-55. Khorevin, V. I. (2008). "The lagena (the third otolith endorgan in vertebrates)." Neurophysiology 40(2): 142-159. Klimley, P. A., S. C. Beavers, T. H. Curtis and S. J. Jorgensen (2002). "Movements and Swimming Behavior of Three Species of Sharks in La Jolla Canyon, California." Environmental Biology of Fishes 63(2): 117-135. Lanca, M. J., M. Machado, C. S. Mateus, M. Lourenco, A. F. Ferreira, B. R. Quintella and P. R. Almeida (2014). "Investigating Population Structure of Sea Lamprey (Petromyzon marinus, L.) in Western Iberian Peninsula Using Morphological Characters and Heart Fatty Acid Signature Analyses." Plos One 9(9): 14. Landreth, H. F. and D. E. Ferguson (1967). "Movements and orientation of the tailed frog, Ascaphus Truei." Herpetologica 23(2): 81-93. Lohmann, K. J., C. M. F. Lohmann and C. S. Endres (2008). "The sensory ecology of ocean navigation." Journal of Experimental Biology 211(11): 1719-1728. Luschi, P. (2013). "Long-Distance Animal Migrations in the Oceanic Environment: Orientation and Navigation Correlates." ISRN Zoology 2013: 23. Luschi, P., S. Åkesson, A. Broderick, F. Glen, B. Godley, F. Papi and G. Hays (2001). "Testing the navigational abilities of ocean migrants: displacement experiments on green sea turtles (Chelonia mydas)." Behavioral Ecology and Sociobiology 50(6): 528-534. 154 Masse, A. K. and C. R. Murthy (1990). "Observations of the Niagara River thermal plume (Lake Ontario, North America)." Journal of Geophysical Research: Oceans 95(C9): 16097-16109. Meckley, T., C. Holbrook, C. Wagner and T. Binder (2014a). "An approach for filtering hyperbolically positioned underwater acoustic telemetry data with position precision estimates." Animal Biotelemetry 2(1): 7. Meckley, T., C. Wagner and E. Gurarie (2014b). "Coastal movements of migrating sea lamprey (Petromyzon marinus) in response to a partial pheromone added to river water: implications for management of invasive populations." Can J Fish Aquat Sci 71(4): 533 - 544. Meckley, T. D., C. M. Wagner and M. A. Luehring (2012). "Field evaluation of larval odor and mixtures of synthetic pheromone components for attracting migrating sea lampreys in rivers." Journal of Chemical Ecology 38(8): 1062-1069. Miller, J. R., P. Y. Siegert, F. A. Amimo and E. D. Walker (2009). "Designation of Chemicals in Terms of the Locomotor Responses They Elicit From Insects: An Update of Dethier et al. (1960)." Journal of Economic Entomology 102(6): 2056-2060. Moore, H. H. and I. P. Schleen (1980). "Changes in spawning runs of Sea Lamprey (Petromyzon-Marinus) in selected streams of Lake Superior after chemical control." Canadian Journal of Fisheries and Aquatic Sciences 37(11): 1851-1860. Mouritsen, H., R. Derbyshire, J. Stalleicken, O. Ø. Mouritsen, B. J. Frost and D. R. Norris (2013). "An experimental displacement and over 50 years of tag-recoveries show that monarch butterflies are not true navigators." Proceedings of the National Academy of Sciences 110(18): 7348-7353. Mueller, T. and W. F. Fagan (2008). "Search and navigation in dynamic environments - from individual behaviors to population distributions." Oikos(117): 654-664. Nathan, R., W. M. Getz, E. Revilla, M. Holyoak, R. Kadmon, D. Saltz and P. E. Smouse (2008). "A movement ecology paradigm for unifying organismal movement research." Proceedings of the National Academy of Sciences 105(49): 19052-19059. Pe'er, G., D. Saltz, H.-H. Thulke and U. Motro (2004). "Response to topography in a hilltopping butterfly and implications for modelling nonrandom dispersal." Animal Behaviour 68(4): 825839. 155 Putman, N. F., E. S. Jenkins, C. G. J. Michielsens and D. L. G. Noakes (2014). Geomagnetic imprinting predicts spatio-temporal variation in homing migration of pink and sockeye salmon. R Development Core Team (2015). R: A language and environment for statistical computing. Vienna, Austria, R Foundation for Statistical Computing. Rodriguez-Munoz, R., J. R. Waldman, C. Grunwald, N. K. Roy and I. Wirgin (2004). "Absence of shared mitochondrial DNA haplotypes between sea lamprey from North American and Spanish rivers." Journal of Fish Biology 64(3): 783-787. Roy, R., J. Beguin, C. Argillier, L. Tissot, F. Smith, S. Smedbol and E. De-Oliveira (2014). "Testing the VEMCO Positioning System: spatial distribution of the probability of location and the positioning error in a reservoir." Anim Biotelm 2: 1. Scapini, F., M. Fallaci and M. C. Mezzetti (1996). "Orientation and migration in Talitrus saltator." Revista Chilena de Historia Natural 69: 553-563. Silva, S., M. J. Araújo, M. Bao, G. Mucientes and F. Cobo (2014). "The haematophagous feeding stage of anadromous populations of sea lamprey Petromyzon marinus: low host selectivity and wide range of habitats." Hydrobiologia 734(1): 187-199. Smith, F. (2013). "Understanding HPE in the VEMCO positioning system (VPS)." Halifax, NS: VEMCO. Sorensen, P. W., J. M. Fine, V. Dvornikovs, C. S. Jeffrey, F. Shao, J. Z. Wang, L. A. Vrieze, K. R. Anderson and T. R. Hoye (2005). "Mixture of new sulfated steroids functions as a migratory pheromone in the sea lamprey." Nature Chemical Biology 1(6): 324-328. Spice, E. K., D. H. Goodman, S. B. Reid and M. F. Docker (2012). "Neither philopatric nor panmictic: microsatellite and mtDNA evidence suggests lack of natal homing but limits to dispersal in Pacific lamprey." Molecular Ecology 21(12): 2916-2930. Taylor, G. K., R. I. Holbrook and T. B. de Perera (2010). Fractional rate of change of swimbladder volume is reliably related to absolute depth during vertical displacements in teleost fish. Teeter, J. H. (1980). "Pheromone communication in sea lampreys (Petromyzon marinus): implications for population management." Canadian Journal of Fisheries and Aquatic Sciences 37: 2123-2132. 156 Ueda, H., M. Kaeriyama, K. Mukasa, A. Urano, H. Kudo, T. Shoji, Y. Tokumitsu, K. Yamauchi and K. Kurihara (1998). "Lacustrine sockeye salmon return straight to their natal area from open water using both visual and olfactory cues." Chemical Senses 23(2): 207-212. Vrieze, L., R. Bergstedt and P. Sorensen (2011). "Olfactory-mediated stream-finding behavior of migratory adult sea lamprey (Petromyzon marinues)." Can J Fish Aquat Sci 68: 523 - 533. Vrieze, L. A., R. Bjerselius and P. W. Sorensen (2010). "Importance of the olfactory sense to migratory sea lampreys Petromyzon marinus seeking riverine spawning habitat." Journal of Fish Biology 76(4): 949-964. Vrieze, L. A. and P. W. Sorensen (2001). "Laboratory assessment of the role of a larval pheromone and natural stream odor in spawning stream localization by migratory sea lamprey (Petromyzon marinus)." Canadian Journal of Fisheries and Aquatic Sciences 58(12): 2374-2385. Wagner, C. M., M. L. Jones, M. B. Twohey and P. W. Sorensen (2006). "A field test verifies that pheromones can be useful for sea lamprey (Petromyzon marinus) control in the Great Lakes." Canadian Journal of Fisheries and Aquatic Sciences 63(3): 475-479. Waldman, J., C. Grunwald and I. Wirgin (2008). Sea lamprey Petromyzon marinus: an exception to the rule of homing in anadromous fishes. Walker, M. M., T. E. Dennis and J. L. Kirschvink (2002). "The magnetic sense and its use in long-distance navigation by animals." Current Opinion in Neurobiology 12(6): 735-744. Wendler, G. and H. Scharstein (1986). "The orientation of grain weevils Sitophilus Granarius influence of spontaneous turning tendencies and of gravitational stimuli " Journal of Comparative Physiology a-Sensory Neural and Behavioral Physiology 159(3): 377-389. Westerberg, H. (1982). "Ultrasonic tracking of Atlantic salmon (Salmo salar L.), 2: Swimming depth and temperature stratification." Report-Institute of Freshwater Research, Drottningholm (Sweden). 157 CHAPTER 4 DOES LARVAL SEA LAMPREY ODOR AID NAVIGATION OR GUIDE HABITAT SELECTION DECISIONS BY ADULTS ABSTRACT The ways in which non-homing fishes, such as sea lamprey (Petromyzon marinus), locate and select reproductive habitat is poorly understood. Prior studies indicated the presence of larval odor in river water increased the likelihood that a sea lamprey entered a river. However, it was not known whether larval odor played a role in navigation (guiding the migrant to the river mouth) or mediated habitat selection by labeling the suitability of a river for spawning. Fixed acoustic telemetry captured sea lamprey movement and a hydrodynamic model coupled with a dye concentration model provided estimates of the Ocqueoc River plume position through time. This allowed for the reconstruction of the hydrodynamic experience of sea lamprey as the approached the coast, encountered river water, and entered a river in the Great Lakes. Once reaching a coast and encountering river water, sea lamprey deflected along the physical coastal edge within the river water labeled region of the lake and approached the river, consistent with a thigmotactic search mechanism rather than a chemotactic search mechanism, where gradients of odorants are important. These observations compliment other evidence to suggest that encounter with river water absent larval odor is sufficient for migrants to transition to local search, while larval odor guides the river entry decision. The river entry decision may have occurred before entry, though it was never reversed following entry. Three synthesized components found in the full larval odor had no apparent effect on the likelihood of river entry when compared to the entry rate previously observed in the Ocquoec River between 1 and 2 years after pesticide based 158 removal of larvae. These known components of larval odor appeared insufficient to bias river entrance; however, the results suggested that the use of the entire larval odor could be a powerful management tool, capable of manipulating the river entry decision. INTRODUCTION Animals can use physical and chemical information to find and select reproductive habitat. Olfactory signals associated with a resource can label directional information that leads to a target and remotely inform the searcher of the profitability of the target (Vickers 2000). When studying animal search, determining whether a particular odor is important to navigation to a target or assessment of target quality is difficult. Sea lamprey adults do not home and rely on the odor of larval sea lamprey from previous generations residing in a stream to select a river (Morman et al. 1980, Teeter 1980, Bergstedt and Seelye 1995, Waldman et al. 2008). Once in a river, sea lamprey are hypothesized to search using odor mediated rheotaxis to locate habitat and mates (Bjerselius et al. 2000, Vrieze et al. 2011, Johnson et al. 2012, Meckley et al. 2014b). However, it is unclear how sea lamprey search in a river plume, defined as the dynamic region of a lake where river water mixes with lake water, or how they enter rivers. Lab tests designed to mimic a river plume demonstrated that sea lamprey are attracted to river water lacking larval odor (Vrieze and Sorensen 2001). This opens the question of whether larval odor or river water plays a role in the decision to enter a river or whether larval odor is important to both river selection and navigation in a river plume. The sea lamprey migration in the lake is just beginning to be described. On the coast in the Great Lakes, sea lamprey display extensive movements parallel to the coast, typified by straight movements while completing vertical excursions from the surface to the bottom of the water column until encountering a river plume (Vrieze et al. 2011, Meckley et al. 2014b), during 159 their non-homing migration (Bergstedt and Seelye 1995, Waldman et al. 2008). The vertical movement is hypothesized to improve encounter with migration cues found in stratified layers of the water column and may be involved in orientation (Vrieze et al. 2011, Meckley et al. 2014b, Chapter 3). Upon encountering a river plume sea lamprey transitioned to intensive movements typified by horizontal turning and continued vertical excursions, as well as regular deflections off the coast in the area of the river plume, as individuals moved through the river plume and localized the river mouth (Meckley et al. 2014b). A general characterization exists of how sea lamprey navigate in rivers and move along the physical coastal edge to encounter river plumes (Vrieze et al. 2011, Meckley et al. 2014b), though how sea lamprey move in river plumes and the functional role of larval odor remains unclear due to the limitations of the tracking equipment of earlier studies. Some have suggested sea lamprey search in river plumes through kinesis (e.g., Vrieze et al. 2011), an undirected search absent directional information, while others have suggested a directed search mechanism reliant on physical edges (e.g., Thigmotaxis, Meckley et al. 2014b). Larval odor could increase selection of a river by labeling good habitat that guides decision making or by playing a role in navigation, allowing sea lamprey to localize the river mouth more easily. Larval odor influences the number of sea lamprey to enter a river, demonstrated by the large reduction in sea lamprey trapped in rivers the year after mass removal of lamprey larvae (Moore and Schleen 1980). Olfaction is thought to be integral to sea lamprey finding and entering rivers in the Great Lakes (Vrieze et al. 2010). Reliance on odor for habitat selection sets up the potential to switch from traditional management (e.g., dams, pesticide) to odor based manipulation of migration routes, guiding sea lamprey into habitat that is easy to manage or is unfit for larval survival (Li et al. 2007). Manipulation of the invasive sea lamprey’s habitat selection with the use of larval odor 160 has been greatly anticipated as an integrated pest management solution, although it requires identification and synthesis of the components in larval odor to be able to activate rivers (Sorensen et al. 2003, Johnson et al. 2006, Hoye et al. 2007). To our knowledge, only three components of the larval odor have been identified and synthesized (Hoye et al. 2007). Field experiments indicate that sea lampreys can detect and respond to incomplete odors to facilitate the search for reproductive opportunities (Fine and Sorensen 2008, Meckley et al. 2014b), but ultimately require encounter with a complete composite of the components that represent a signal prior to selecting spawning habitat (Wagner et al. 2009, Meckley et al. 2012, Meckley et al. 2014a) or a mate (Johnson et al. 2006, Johnson et al. 2009, Luehring et al. 2011). The river mouth would be the most likely location for a partial odor to inform a decision, as the complete odor could be the most dilute rendering the most concentrated components the most reliable indicators of larval presence (Meckley et al. 2012). If a collection of partial odors fails to elicit entry at the river mouth, it would indicate the need for identification and synthesis of additional compounds found in larval odor. Here, we report a field experiment that evaluated whether larval odor influences the riverentry decision or whether the cue is important to navigation. Secondarily we provided a test of whether the addition of synthetic larval odor increases entry beyond what is expected two years after TFM treatment. During a two-year study in 2010 and 2011, we documented the three dimensional movements of acoustically tagged sea lamprey as they approached, entered, or bypassed a river under one of two conditions: (1) low larval odor, following larval removal with a pesticide (TFM), below (2008) and above a barrier (2009) (2010); and, (2) higher larval odor, created by larval recruitment, plus synthetic larval odor components Petromyzonamine disulfate (PADS), petromyzosterol disulfate (PSDS), and petromyzonol sulfate (PZS) to a 1 x 10 -12 M 161 concentration (2011). A coupled hydrodynamic model and dye concentration model provided environmental information including water current direction and speed, water temperature, and the presence of river water at each fish position for the entire study period allowing us to characterize the sea lamprey migration experience with unprecedented precision. We specifically tested three hypotheses. First, sea lamprey will encounter the river plume at the same rate between years, although more will enter in 2011 due to a higher larval odor concentration a year after any TFM treatment in the Ocqueoc River and following natural recruitment of larvae (H1). Second, higher larval odor will not increase how close sea lamprey come to the river mouth but will influence how sea lamprey move in the river plume for those that encounter river water (H2). Third, the addition of the known synthetic odor components of the full larval odor composition will be insufficient to increase entry rate beyond the expected river entry rate (H3). The expected river entry rate was based on the increase historically observed between one and two years after pesticide treatment in the Ocqueoc River (H3). METHODS General Methods Acoustically tagged sea lamprey were released at three staggered locations 700 m north of the Ocquoec River (45.490278°, −84.072931°) in Hammond Bay (Lake Huron) at the center of a VEMCO positioning array capable of providing 3-D positions every 30 seconds ( Figure 4.1). Sea lampreys were released over two different spawning seasons (N=72, 2010; N=79, 2011) under two conditions including an assumed lower (2010) and higher (2011) larval population condition. Prior to the first year of the study (April 2010) both the lower (6 October 2008) and upper (22 August 2009) sections of the Ocqueoc river, separated by a dam, were treated with the pesticide 3-trifluoromethyl-4-nitrophenol (TFM). TFM treatment vastly reduced 162 the population in the Ocquoec River, although the Ocquoec River still receives a reduced but reliable sea lamprey run the year after treatment, as treatment is required every four years prior to larvae transforming to parasites and out migrating from the river to the lake (Aaron Jubar, US Fish and Wildlife Service, personal communication). Study Site The Ocqueoc river had an annual spring median discharge during trials of 2.64 m3·s-1 in 2010 and 3.9 m3·s-1 in 2011 (range: 0.6-9.7 m3·s-1 2010, 1.8-8.6m3·s-1 2011). A telemetry system (VPS, VEMCO) with a similar configuration of acoustic receivers (VR2W) and coverage in 2010 and 2011 was composed of 41 receivers located in a 0.90 km2 in 2010 and 37 receivers covering 1.57 km2 in 2011 and centered on the mouth of the Ocqueoc River (45.490278°, −84.072931°). The receiver configuration was in diamond formations with between receiver spacing ranging from 75 m to 250 m (Figure 4.1). The depth of the site ranged from 0 to 5.8 m within the array but sea lamprey were positioned at greater depths on the fringes of the array. Nine independently moored synchronization transmitters (VEMCO model V16-2H, 69 kHz) were deployed in stationary positions through the array. In addition to the receivers in the acoustic array, receivers were positioned in the Ocqueoc River at six locations monitoring upstream progress in 2010 and 2011. Experimental Subjects The ethical treatment and acoustic tagging of sea lamprey was approved by the Michigan State University Institutional Animal Use and Care Committee via animal use permit 02/10-02000. Adult female sea lamprey were obtained from the Cheboygan River at a barrier integrated trap (2010: N=72, trapped from 15, April to 10, May; 2011: N=79, trapped from 30, April to 15, May). In 2010, Sea lampreys were 417-567 mm in length (mean 495 mm) and weighed 148-409 163 g (mean 247.7 g). In 2011, Sea lampreys were 458-588 mm in length (mean 506 mm) and weighed 191-389 g (mean 274 g). All Sea lamprey were held in 150 L flow through tanks that cycled ambient Lake Huron water (100% exchange every 2 h) and experienced a natural light cycle. Acoustic transmitters transmitted signals every 15-45 s, though one contained a pressure sensor that transmitted pressure every transmission (2011) and the other transmitted pressure every other transmission (2010) (model V9P-2H,Vemco, Halifax, Nova Scotia, Canada 9mm D x 47 mm L, mass: 6.4 g in air, 3.5 g in water, power output 150 dB (re 1 µPa at 1 m)) (model V92H,Vemco, Halifax, Nova Scotia, Canada, 9mm diameter, 29 mm length: mass: 3.6 g in air, 2.2 g in water, power output 151 dB (re 1 µPa at 1 m)). The model with alternating pressure transmissions was selected in 2010 due to its shorter duration transmission length to ensure transmission collisions did not prevent position acquisition. Acoustic tagging procedures followed methods by Meckley et al. 2014b. Prior to surgery sea lamprey were anesthetized by immersion in 0.2 mL·L−1. The anesthetic solution used in this experiment was composed of 2 mL of clove oil (minimum 84%-88% eugenol, Lot No. HB9387, Hilltech Canada Inc. Vankleak Hill. Ontario, Canada) diluted in 18 mL of 70% ethanol and vigorously mixed into 10 L of Lake Huron water. Sea lamprey took 634 ± 11 s to reach stage four of anesthesia, as denoted by unresponsiveness to touch but retained gill movement, at a temperature of 8.0 ± 0.1 O C (mean ± 1 SE). The surgical procedure took an average of 319 ± 7 s (mean ± 1 SE) and equilibrium was regained in 527 ± 32 s (mean ± 1 SE). Transmitter-implanted subjects were held for 72 h prior to release to ensure metabolism of stress compounds (Close et al. 2003). Pheromone application Synthesized PADS (MW: 722 g), PSDS (MW: 625 g), and PZS (MW: 473.4 g) were applied continuously at a ratio of 1:1:1 based on molecular weight, from 4 May to 6 June 2011, 164 900 m from the river mouth to the middle of the water column in the center of the stream. PADS, PSDS and PZS were synthesized by Bridge Organics, Inc. (Kalamazoo, Michigan, USA). The purity of each product was confirmed at >95% via high-pressure liquid chromatography (HPLC) with a purity evaporative light scattering detector and mass spectrometry by Xiodan Xi at Michigan State University. Pheromone applications were permitted by the US Environmental Protection Agency through experimental use permit 75437-EUP-1 and the Michigan Department of Environmental Quality pursuant to Rule 97. Prior to application, we created 1.0 mg·mL−1 stock solutions for each compound via dissolution into 1:1 v/v methanol:water. A single stock solution for each compound was used during the course of the study. The final odor was prepared by mixing the required amount of each stock solution into a necessary amount of river water before being pumped into the stream at a fixed rate of 600 mL·h−1 via a battery-operated programmable peristaltic pump (Admiral Reef Dosing Pump, Norwich, Connecticut, USA). VPS Array Performance and data treatment The performance of the VPS array in terms of position accuracy was tested by comparing the VPS position estimates of V9P-2H transmitter to GPS measured positions (Trimble Geo XH, post processed). Array performance was tested by either attaching a transmitter to an anchored line as described in Meckley et al. 2014a (stationary testing) that provided longer term information (2010: Location 1: 17 June 2010 19:43 01 July 2010 14:34 (N= 29,355 detections), Location 2: 01 July 2010 14:52 to 08 July 2010 17:03 (N= 16,400 detections); 2011: 09 June 2011 016:30 to 24 June 2011 04:07 (N= 26,628 detections)) or by dragging transmitters while inside the body cavity of a sea lamprey that was recently deceased, located 1 m off the bottom and below a floating boat powered by an electric motor with the GPS mounted directly above the tag monitoring the tags’ true position (2010: N=138; 2011: N=309). The mean unfiltered 165 accuracy estimates from stationary tests were 2.0 ± 6.6 m (2010 Test 1), 1.1 ± 5.3 m (2010 Test 2), 1.5 ± 1.4 m (2011 Test 1) (mean ± SD). The mean unfiltered accuracy from the mobile tests was 3.0 ± 6.2 m (2010) and 5.4 ± 12.5 m (2011) (mean ± SD). Three data quality objectives were selected, including the need to ensure that trajectory data (e.g. turn angles) could be considered, designations of when the sea lamprey were stopped or moving were reliable based on first passage time; large losses in data did not occur while achieving the first two objectives that could result in misrepresenting the tracks (Table 4.1). Trajectory based analyses often break down when average position error is > 10 % of the mean step length (Bradshaw et al. 2007). In this case the mean step length was 17.7 m, resulting in a filter objective of less than 1.77 m average error (Meckley et al. 2014 a; Table 4.2). To avoid incorrect behavioral assignment of swimming or stopped we wanted to remove large erroneous positions (e.g. >15 m (citation)). Finally we wanted to maximize the removal of inaccurate data and retention of accurate data to ensure proper assignment of when the fish encountered river water in our dye concentration model, a step that should be possible in our array conditions, this meant retention of 95 % of acceptably accurate (<6 m) data and removal of 99 % of inaccurate data (> 6 m). If a large amount of data was lost, the analysis that required the data quality objective would have been replaced. An HPE filter of 8 was selected in 2010 and an HPE filter of 7 was selected in 2011 based on the three criteria regarding the position precision estimates and data retention (Table 4.1, Table 4.2). The filtered coverage was 8.22 km2 in 2010 and 10.95 km2 in 2011. The filter removed 9% of fish positions inside the array in 2010 and 7% of positions inside the array in 2011. Hydrological Data Bathymetry in the region of the array was recorded via 50 meter transects oriented 166 towards and away from shore and then repeated parallel to shore. Bathymetric data were collected with an acoustic Doppler current profiler (ADCP; Workhorse Rio Grande 1200 kHz, Teledyne RD Instruments, Poway, CA). Data from four ADCP beams were combined in real time with geographic locations from a WAAS-enabled GPS (model Geo-XT, Trimble Navigation Ltd., Sunnyvale, CA) using the software program BathMapper (courtesy of A. Blake, U.S. Geological Survey, California Water Science Center). Total water column depths at fish positions were interpolated from the georeferenced depth data using inverse distance weighting. A 2 MHz upward facing Aquadopp Current Profiler (ADCP, Nortek-USA) was deployed on the bottom in 7.9 m of water at two nearby locations asynchronously from the 8 May to 30 May (45.50207737, -84.07552895) and from the 30 May to 17 June, 2011 (45.50207846, 84.07552753). The ADCP recorded water movement in 3 dimensions at twenty slices through the water column and was used to check the hydrodynamic model (see below for more information). Five HOBO temperature loggers were deployed each at three locations in front of the river mouth in the lake (from mouth; left: 300 m, center: 150 m, right: 350 m) farther offshore (700 m) and in the river for the duration that receivers were deployed. We estimated the stream discharge daily at the application site using the midsection transect method with a Doppler flow meter (Flo-Mate Model 2000, Marsh-McBirney)(Gore 2007). Hydrodynamic Models: Two models describe the circulation of lake water in Hammond Bay of Lake Huron (Numerical Model for 2010 and 2011) and how river water mixes with lake water forming a zone of river influenced lake water, which we refer to as the river plume (dye concentration model). We estimated the water conditions at each fish position and through the region including current speed and direction, water temperature, and dye concentration as an estimate of the presence of river water. 167 Numerical Model: The three-dimensional unstructured grid numerical model (FVCOM; minimum grid size: 9.15 m2) is used to describe circulation and thermal structure in Hammond Bay, Lake Huron (Nguyen et al. 2014) (Figure 4.2). The model solves the hydrodynamic primitive equations using the hydrostatic assumption in the vertical direction with the Boussinesq simplification for convective flows. The continuity, momentum, and temperature equations are shown in Equation 1 to Equation 5. 𝜕𝑢 𝜕𝑣 𝜕𝑤 + + =0 𝜕𝑥 𝜕𝑦 𝜕𝑧 (1) 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 1 𝜕𝑃 𝜕 𝜕𝑢 +𝑢 +𝑣 +𝑤 − 𝑓𝑣 = − + (𝐾𝑀 ) + 𝐹𝑢 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌0 𝜕𝑥 𝜕𝑧 𝜕𝑧 (2) 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 1 𝜕𝑃 𝜕 𝜕𝑣 +𝑢 +𝑣 +𝑤 + 𝑓𝑢 = − + (𝐾𝑀 ) + 𝐹𝑣 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜌0 𝜕𝑦 𝜕𝑧 𝜕𝑧 (3) 𝜕𝑃 = −𝜌𝑔 𝜕𝑧 𝜕𝑇 𝜕𝑇 𝜕𝑇 𝜕𝑇 𝜕 𝜕𝑇 +𝑢 +𝑣 +𝑤 = (𝐾𝐻 ) + 𝐹𝑇 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑧 (4) (5) Here (𝑢, 𝑣, 𝑤) are velocity components in horizontal and vertical directions, respectively; 𝜌 is the density; 𝜌0 is the reference density; 𝑇 is the temperature; 𝑃 is the pressure; 𝑓 is the Coriollis parameter;𝑔 is the acceleration due to gravity; (𝐹𝑢 , 𝐹𝑣 , 𝐹𝑇 ) are the horizontal momentum and thermal diffusion terms. Vertical eddy viscosity and diffusivity(𝐾𝑀 , 𝐾𝐻 )are modeled using the Mellor-Yamada 2.5 level turbulence closure scheme (Mellor and Yamada 1982, Galperin et al. 1988). The horizontal diffusion coefficients are calculated using the Smagorinsky turbulence closure model (Smagorinsky 1963). Comparisons between vertically averaged current and temperature from the ADCP observations and simulations provided a good fit suggesting the 168 models effectively predicted water conditions in Hammon Bay (Nguyen et al. 2014) (Figure 4.3, Figure 4.4). Dye Concentration Model; the dye concentration model is coupled with the hydrodynamic model and was solved using the following equations: 𝝏𝑫𝑪 𝝏𝒕 + 𝝏𝑫𝒖𝑪 𝝏𝒙 + 𝝏𝑫𝒗𝑪 𝝏𝒚 + 𝝏𝒘𝑪 𝝏𝝈 𝟏 𝝏 𝝏𝑪 − 𝑫 𝝏𝝈 (𝑲𝒉 𝝏𝝈) − 𝑫𝑭𝒄 = 𝑫𝑪𝟎 (𝒙, 𝒚, 𝝈, 𝒕) where C is the concentration of the dye, D is the total depth, u, v, and w are the x,y, and  components of the water velocity, Kh is the vertical diffusion coefficient, Fc is the horizontal diffusion term, and C0 is the initial dye concentration (Chen et al., [2012]).The dye concentration is calculated at every grid cell throughout all layers in the water column in which the initial condition of dye concentration is zero at every grid cell of the FVCOM grid. The dye concentration from river was kept at 100ppm throughout the simulation time though the amount of water released is based on river discharge and influences C0. The river plume is then determined based on the values of the vertically-integrated concentration at every grid cell. The comparisons for velocity and temperature from hydrodynamic model are acceptable, providing confidence for the dye concentration model as it is coupled with hydrodynamic model. The fact that this model is coupled with the hydrodynamic models and those models have a good fit with observed data, we have confidence in the predictions of the dye concentration model (Figure 4.3, Figure 4.4). Identifying active sea lamprey To characterize the periods of sea lamprey activity in the lake, we first had to robustly classify sea lamprey positions as active or stopped. We achieved this with a first passage time classification method performed in R (R Development Core Team 2015). The first passage time 169 tool classified a position as moving if it left a radius of 10 meters in 250 seconds, had a minimum displacement of 15 m in a 3 position moving average, and at least 4 consecutive moving observations. The characterization method was checked through visual inspection and was robust to imprecision in acoustic positioning after filtering (Gurarie et al. 2015). Each night of activity that an individual sea lamprey was observed in the array was characterized into a unique bout of activity for that individual. For each step comprised of two consecutive positions that was separated by less than ten minutes was used to calculate trajectory data. The midpoint of each step was used to estimate environmental information experienced during the step and the distance to key features (e.g. nearest coast or river). The trajectory data included heading (Phi; radians 0-2Π), turn angle (Theta, phase angle; radians Π (right turn) – 0 (straight) - - Π (left turn), and ground speed (m/s). The lake water temperature, lake water current magnitude and direction, whether during a bout a sea lamprey had yet encountered river water, the distance to the nearest coast and the distance to the river were all determined in R for each mid-point. Characterizing river plume encounter and entry We separated individuals into two categories including individuals that stopped in the array after release (A) and all individuals released (B), as sea lamprey that stopped in the array after release provided a more complete depiction of the start of the lake migration. Both categories were further separated into sea lamprey that encountered river water on the first night and those that did not encounter river water. To determine if differences occurred in the rate of entry between years on the first night or ever, the rates at which sea lamprey encountered river water and entered the Ocqueoc River for both groups (A and B) were evaluated with a chi square tests of independence (χ2)(R Development Core Team 2015). To consider if the rate of river 170 entry changed when individuals encountered the river plume on more than one night, entry of all sea lamprey (B) were tabulated in light of the number of nights they encountered the river plume before entry (rate of all available). Similarly the specific entry rates of individuals that returned and encountered river water on 1, 2, or more than 3 nights were quantified. To evaluate the effect of a host of variables on river entry, a mixed effects logistic regression with a binary response variable (river entry: 1 or 0) was implemented in R using glmer (lme4 package, Bates et al. 2014). The binary response variable river entry (E) was considered for each night of activity that a sea lamprey encountered river water based on the dye model (dye concentration > 5); where individual (ID) was a random effect since multiple nights of activity could occur for the same sea lamprey. The fixed effects included year (Y), iterative number of nights encountering river water (N), a year by night number interaction(YxN), average nightly discharge of the Ocqueoc River (DC: m3·S-1), average lake temperature encountered (LT: C ̊), average nightly river temperature (RT; C ̊ ), the difference in experienced lake temperature and river temperature(DFT; C ̊ ), minimum distanced reached from the river mouth (D: km), and average water velocity experienced in the lake (V) (Table 4.3). We used AIC (Akaike’s information criterion) to guide model selection (Burnham and Anderson 1998; Burnham and Anderson 2002). We used AIC without criteria adjusted for sample sizes (AICC), as AICC requires a known number of independent observations and this is difficult to determine in mixed-effects models (Dodge et al. 2014; Burnham and Anderson 1998). All analyses were performed in R (R Development Core Team 2015), including the lme4 package. To determine what influenced the way sea lamprey move we quantified movement behavior with ground speed (Vm, m·s-1) and path straightness. Activity was considered on the first night of activity, for active sea lamprey that stopped in the array following release and 171 encountered river water on the first night. Each step was associated with a ground speed (V) and all location associated variables were based on the mid-point of the step. Path straightness (S) was estimated by looking at the total path length versus the beeline distance for sets of ten consecutive points. We fit models with a nonlinear mixed-effects model in the formulation described by Lindstrom and Bates (1990), using the function “nmle” in R (R Development Core Team 2015). This method allowed for incorporation of the random effect of each individual lamprey (ID) and first order auto-correlation in the residuals to be accounted for in the models. The significance of effects within the best model was assessed by evaluating the t-values and pvalues; where large t values and p-values less than 0.05 were considered significant. The closest distance each sea lamprey was observed to the river for all sea lamprey and for those that encountered river water and did not enter was determined and tested for differences between years with a Mann-Whitney U test with an alternative hypothesis that there was no difference between the means. The time it took for entry to occur was also tested between years with a Mann-Whitney U test that sea lamprey would take longer to enter in 2010. The upstream progress of sea lamprey was monitored at 3 pairs (N=6) of acoustic receivers in the river that could capture passage and allow for interpretation of upstream or downstream movement. We used these receivers to determine river entry and the time it took for sea lamprey to reach the lamprey barrier on the Ocqueoc River located 7916 m from the river mouth. Movement Patterns Ground speed was evaluated as a continuous response variable with respect to year (Y), water temperature at fish location (T, C ̊ ), an interaction between current magnitude and the angle between fish heading and current direction (as a turn angle in radians: 0-Π, where 0 indicated movement into the current and Π away from current) (CxF, radians 0- Π), whether the 172 fish position was inside or outside of the river plume (dye concentration of > 5) (P), and whether encounter with river water occurred yet in each path (E). The average time difference between points was 49 s with a max of 569 s (<10 min). We included the cumulative distance traveled (D) for each sea lamprey path in each model as previous observations captured increasing ground speed at the start of the night (Chapter 3). Animal identification number (ID) was included as a random effect and first order autocorrelation in the residuals was taken into account. Ten competing mixed effects models were fit (Table 4.4). Path straightness (Sm) ranges from 0-1 and was evaluated as a continuous response variable with respect to the fixed effects of whether sea lamprey had encountered river water in the path yet (E), whether sea lamprey were in the river plume (P), year (Y), total cumulative path length (D), and additional interactions within these variables. Changes in path straightness are often used to identify local search behavior (Lohmann et al. 2008, Dodge et al. 2014). Animal identification number (ID) was included as a random effect and first order auto correlation in the residuals was taken into account. Ten competing mixed effects models were fit (Table 4.5). To characterize how sea lamprey migrate outside of the river plume, we summarized path length versus the approach ratio (distance moved towards the coast/total distance moved) to the coast where -1 would indicate progress exactly away from the coast, 0 would be movement parallel to the coast and 1 would represent movement towards the coast. The distance from shore when sea lamprey exited the array while moving parallel to the coast was summarized and the direction chosen with respect to the local current encountered was tabulated. To test if the direction chosen was opposite the local water current encountered while moving in the array, we tested the proportion of sea lamprey to move east on a night when the current was moving from the east or from the west with a chi square test (χ2). 173 Effect of Synthetic Odor The effect of synthetic larval odor is difficult to evaluate in light of the presence of larvae in the system. The only metric for evaluating the influence of the three synthetic larval odor components comes through comparison of the historical entry rate of wild sea lamprey one and two years post TFM treatment and compared to the entry rate of the sea lamprey released in our study. The percentage of sea lamprey to enter the Ocqueoc River of the total available migratory sea lamprey in Lake Huron was achieved by comparing population estimates provided by Jessica Barber of the Fisheries and Wildlife Service. The Ocquoec River population estimate was based on mark recapture estimates from the Ocqueoc River alone and the Lake Huron estimate was estimated based on extrapolating mark recapture data from a set of reference streams (Mullett et al. 2003. Although there are uncertainties in comparing the magnitudes of estimates through time, the general pattern post pesticide application is robust. RESULTS Sea Lamprey Activity Sea lamprey primarily started moving during the half hour before nautical twilight and stopped in the half hour before sunrise (Figure 4.5A). River entry was occasionally observed prior to nautical twilight (set), although river entry was never observed after nautical twilight (rise) in the morning (Figure 4.5D), despite activity in the lake extending until and occasionally after sunrise (Figure 4.5B). There was some evidence of multiple pulses of river entry, including a pulse of entry after nautical twilight (set), 2.5 hours after nautical twilight (set), and just prior to nautical twilight (rise) (Figure 4.5C, Figure 4.5D). Stopping in the array during nautical twilight was rare and prolonged periods of inactivity consistent with staging were not observed. 174 Characterizing river plume encounter and river entry There was no difference in river plume encounter between 2010 and 2011, although all measures suggested a higher rate of entry in 2011, in support of hypothesis 1. The majority of sea lamprey stopped in the array after release in 2010 (53 of 72; 74%) and 2011 (59 of 79; 75%). A low proportion of sea lamprey that exited the array coverage immediately after release were not observed again (N=9, 2010; N=11, 2011). Our observations showed no difference in the likelihood of encountering river water on the first night in 2010 (31/53; 58%) or 2011 (41/59; 69%) for trial animals that stopped in the array after release (χ2 = 1.03, df=1, p=0.3; R-core Team 2015). Similarly, there was no significant difference in the likelihood of ever encountering river water from the Ocqueoc River for all released in 2010 (60/72; 83%) and 2011 (61/79; 77%) (χ2 = 0.54, df=1, p=0.46); which includes individuals that encountered river water while moving after release, and those that encountered river water after the first night. Of those we observed moving at night, 48 (66%; 2010) and 53 (67%: 2011) encountered river water during search, the eventual river entry rate of these individuals was 0.5 in 2010 and 0.85 in 2011. Sea lamprey entered the Ocqueoc river at a higher rate (1.72) in 2011 (45/79; 57%) than in 2010 (24/72; 33%) (χ2 = 7.55, df=1, p=0.006). Sea lamprey that stopped in the array and encountered the Ocquoec River water on the first night in 2011 were significantly more likely to enter the river on the first night than in 2010 (2010: 7/31, 23%; 2011: 18/36, 50%; χ2 = 0.18, df=1, p=0.04). Sea lamprey that ever encountered river water during the year eventually entered the Ocqueoc River at a higher rate (1.85) in 2011 (45/61; 74%) than 2010 (24/60; 40%) (χ2 = 12.732, df=1, p<0.001). The pattern of higher entry in 2011 than 2010 was consistent for those that did not enter on the first night and were observed encountering river water on additional nights (Figure 4.6). 175 The best fitting models for river entry included year and minimum distance reached from the river, and either the lake temperature experienced (Model 10) or the difference in lake temperature experienced and the river temperature (Model 11) (Table 4.3). Only year (year 2011; β=0.89), and the effect of the minimum river distance was significant (-2.1), where minimum distance observed related to a reduced rate of entry (p<0.5) (Table 4.6, Model 11 output). River entry did not appear to be influenced by discharge or absolute river temperature. The number of nights observed in the array did not appear as an important factor as we expected, though the pattern of sea lamprey entering if they returned in 2011, and not entering upon return in 2010 was present in the interaction effect between night and year (Figure 4.6). The distance sea lamprey were observed from the Ocquoec River mouth did not vary between 2010 and 2011 for all sea lamprey, in support of hypothesis 2 (mean ± 1 SD: 210 ± 266 m (2010); 129 ± 177 m (2011; Mann-Whitney U = 1065, p=0.54) or for those sea lamprey that didn’t enter the river (mean ± 1 SD: 291 ± 281 m, 2010; 349 ± 236 m, 2011) following river water encounter (Mann-Whitney U =229, p=0.25) (Function “wilcox.test”; R Development Core Team 2015). Sea lamprey did take significantly longer to enter the Ocqueoc River in 2010 (116.9 ± 125 hours; mean ± 1 SD) than in 2011 (86.8 ± 142 hours) based on a one sided Mann-Whitney U with an alternative hypothesis that sea lamprey would enter more quickly in 2011 (W=1006, p=0.046) (Function “wilcox.test”; R Development Core Team 2015). In 2010, 24 of 72 sea lamprey (33%) entered the Ocqueoc River and were detected passing a receiver located 342 m from the river mouth and never reversed course past that point until late in the year once spawning had occurred. Sea lamprey that moved downstream were likely dead or dying, as determined by tags that were stationary in the array and based on recovery of some dead individuals. In addition, two sea lamprey moved close enough to the river 176 mouth or briefly entered the main channel and were detected by a receiver located 103 m from the river mouth. In 2011, 45 of 79 (57%) sea lamprey entered the Ocqueoc River and were detected passing a receiver located 342 m from the river mouth, and never reversed course and exited the river. One sea lamprey was detected on the lowest receiver and did not continue upstream. Two non-trial subjects (2 of 12) released prior to the start of odor pumping in 2011 in an attempt to capture staging behavior, did enter the river and go within the detection range of the receiver located 342 meters upstream before exiting the river, with one of the two returning to the river and progressing upstream in the same night. These two individuals were released in the visible river plume. In 2010, 19 of 24 sea lamprey reached the upper receiver, including 4 that reached the upper receiver on the first night, equaling an upstream ascent of 0.44 ± 0.11 m·s1 (mean ± 1 SE) over the 7.9 km stretch of river. In 2011, 41 of 45 sea lamprey reached the upper receiver including 5 on the first night that had a ground speed of 0.38 ± 0.08 m·s-1 (mean ± 1 SE). The average time to reach the upper receivers (5 or 6) after reaching the second receiver was 44.8 ± 14.6 hours (2010) and 35.5 ± 6.8 hours (2011) (mean ± 1 SE). The time to reach the upper receivers varied from 3.9 – 290.5 hours in 2010 and 4.1 – 187 hours in 2011. Movement Patterns Ground speed did not vary between 2010 and 2011, in support of hypothesis 2 (body length/Second; BL·S-1). The model comparison suggested that Model 9 had the best fit. The best model included temperature (T), if the sea lamprey was in the river plume (P), whether river plume encounter occurred (E), the interaction between lake current magnitude and whether the fish was moving towards it or away from it (CxF), and the interaction effect of cumulative path length and before and after encounter. Year had no apparent effect, as the two worst models included year (Table 4.4). The Model 9 fixed effects parameter estimates revealed that ground 177 speed increased with warmer water temperature (β=0.02, p < 0.05), increased following river plume encounter (β=0.36, p < 0.05), was influenced by the current direction fish movement interaction (β=0.29, p < 0.05), and prior to river encounter speed increased with cumulative path length (β=0.16, p < 0.05) but not after encounter (β=-0.001, p=0.66). Despite being part of the best model the effect of being in the river plume was small (β=0.002) and insignificant (p = 0.88) (Table 4.7, model 9 output). Similarly, path straightness did not vary between 2010 and 2011, in support of hypothesis 2. The model comparison suggested that Model 9 and Model 6 were the best fit. The best two models included whether the sea lamprey was in the plume (P) and Model 9 included the interaction effect between cumulative distance traveled (D) and whether the river plume had been encountered (E) (Table 4.5). Model 9 fixed effects parameter estimates indicated that paths were more sinuous in the plume (β= -0.0045, p < 0.05), and that the path became straighter after settlement and before encounter with the plume as the path progressed (β= 0.048, p < 0.05), and again became straighter after encounter as the path progressed but the effect was weaker (β= 0.018, p < 0.05) (Table 4.8; model output). Sea lamprey that did not encounter river water (N=33) in the array on the first night moved in a circuitous path before approaching the coast and then turning and moving parallel to the coast at some distance (Figure 4.7, Figure 4.8). The average distance from the coast at the final point exiting the array when moving parallel to the coast was 303 ± 221 m. More sea lamprey exited the array to the west (N=17) than to the east (N=7), while 9 were last observed moving to the north or south (χ2 = 4.17, df=1, p< 0.05). Individuals were not more likely to move east when current came from the east rather than the west (0.38 moved east with east current, 178 N=8; 0.15 moved east with west current, N=13) (χ2 = 0.39, df=1, p=0.735, R function “prop.test”). Effect of synthetic odor An annual increase in the number of sea lamprey estimated to enter the Ocquoec River of the entire Lake Huron population is observed. Rates increased on the order of 1.8 (1995-1996), 1.3 (1999-2000), and 3.1 (2003-2004), an average historical rate increase of 2.1 times from one to two years after treatment. The observed rate increase of untagged sea lamprey was 2.1 for 2010 to 2011 for the Ocqueoc River population estimate for our two-year study, matching the historical average. To test if our observed total entry rate of sea lamprey that stop in the array on the first night and encounter river water was higher than the expected 2.1 historical average increase in entry, we ran a chi square test with an alternative hypothesis that the rate observed in 2010 of 0.35 would rise to greater than the expected 0.71 rate in 2011. 41 sea lamprey stopped in the array after release and encountered river water with 28 eventually entering the river. We would expect significantly more than a 29.11 individuals to enter if the entry increased with the average increase between year one and year two. There was no evidence of a greater rate of entry than would be expected two years after TFM treatment, which suggested no effect of synthetic larval odor (hypothesis 3)(χ2 = 0.0, df=1, p=0.50, alternative H0: “greater than”; function prop.test). DISCUSSION These results support our hypothesis that sea lamprey enter rivers with higher larval sea lamprey odor at a higher rate (H1). The effect of larval odor is consistent with altering the decision of sea lamprey to enter a river but does not have a measurable influence on how sea lamprey navigated in the river plume, as ground speed, path straightness, and the minimum 179 distance sea lamprey were observed from the river mouth was not influenced by the effect of year (H2). No evidence was obtained supporting an effect of synthetic larval odor on river entry (H3). We characterized the sea lamprey migration and report which aspects of the prior descriptions of sea lamprey migration in the lake are consistent with our observations. We observed four general patterns of movement. First sea lamprey performed maneuvers consistent with the offshore pattern described as an orientation mechanism (Chapter 3), in which sea lamprey turn in a persistent direction and move slowly before approaching the coast at a faster ground speed and with a relatively straight course (Figure 4.9). Second, if sea lamprey did not encounter river water at some point when approaching the coast (mean 303 m from shore), they turned and moved parallel to the coast (Figure 4.8) more often to the west (Figure 4.7), with straight movements, presumably to encounter a river plume (Vrieze et al. 2011, Meckley et al. 2014b). The dominant current of Lake Huron would come from the west (Beletsky et al. 1999), though the local current experienced by individuals did not relate to the direction individuals exited the array (Figure 4.7). Third, when sea lamprey did encounter river water most individuals made repeated passes up and down the coast in the river plume, which switching between movements towards and away from the coast in an apparent rebounding pattern and counter turning at presumably the river plume edge in a behavior we term, coastal rebounding that was associated with faster ground speed and a more sinuous path. A series of tight rebounds were often observed before sea lamprey entered the river and we postulate with others (e.g., Vrieze et al. 2011, Johnson et al. 2012, Meckley et al. 2014b) that this is the point where sea lamprey transition from edge guided search to odor mediated rheotaxis. 180 Movement outside of the river plume After stopping on the bottom in the array for the day most sea lamprey began moving shortly after nautical twilight, those that did not encounter river water approached the coast and then turned and moved parallel to the coast while either making regular vertical movements between the surface and bottom, alternating between periods at the surface and periods at the bottom, or by moving primarily on the bottom. The majority of sea lamprey exited the array towards the west (55% west, 25% north, 21% east), consistent with observations that most sea lamprey released 3.3 km from shore in Hammond Bay went west if they didn’t enter the closest river and were observed at least 10 km away from the release point (60% (west), 20% (east), 20% (north) (Chapter 3). No advantage to moving in a consistent direction was identified unless a heading in a particular direction allows for the movement parallel to the coast to be maintained more easily or a direction is energetically more efficient. There was no apparent influence of the local current encountered on the direction sea lamprey moved along the coast (Figure 4.7), although the dominant lake current in this area of the lake would be from the west to the east (Beletsky et al. 1999). It is unclear if more sea lamprey moved west as a result of an inherent bias associated with handedness (Kells and Goulson 2001) or if it is due to a geographic or hydrological feature that is valuable to navigation. Observations of the sea lamprey migration on a coast oriented in a different direction would be valuable, as all observations of the lake migration to our knowledge have occurred in or near Hammond Bay, Michigan. It is important to note that eight of the 33 sea lamprey that did not encounter river water and exited the array on the first night returned and entered the river on a subsequent night, clouding the picture of whether sea lamprey move in a consistent compass direction or what features guide the direction of search along the coast. 181 Sea lamprey are likely capable of relying on multiple features of the coastal environmental as they traverse the coast in search of river water, as redundancy in search mechanisms is common (Able 1991). The same environmental features including the bathymetric slope that sea lamprey were postulated to use to return to a coast in chapter 3 could allow sea lamprey to move parallel to the coast; including, water column depth, absolute hydrostatic pressure, or water current. In addition to these feature, physical features of the lake bottom such as the coastal wave zone comprised of a comparatively steep series of banks and troughs that run horizontal to the coast and are easily observed on satellite images could be valuable (Figure 4.10). Use of the bottom structure could explain the periods of movement observed on the bottom before or after periods of vertical casting. A thermal bar along the coast also forms in spring when water above and below 4 degrees Celsius mixes, creating a physically distinct water mass along the coast (2-5 km wide, Rao and Schwab 2007), though this zone should extend beyond the distance sea lamprey were observed moving parallel to shore it still could label the coastal zone. Although blinded sea lamprey were capable of completing a river migration (Binder and McDonald 2007), vision may be valuable to the nocturnal migration as sea lamprey’s eyes are sensitive to light, although it is unclear if this equates to strong night vision (Morshedian and Fain 2015). Our observations confirm the pattern of extensive search parallel to the coast but do not confirm a mechanism. River plume encounter River entry hinges on sea lamprey first encountering river water in the lake, localizing the river mouth, and finally entering the river. The number of sea lamprey to encounter a river plume will depend on both the rate of encounter and the number of sea lamprey to move past the river mouth. Olfaction is postulated as the primary mechanism for detecting river plume encounter; 182 however, other factors such as temperature or anomalies in river current at the mouth could be useful and could explain how nasally occluded sea lamprey still entered rivers in the Great Lakes (Vrieze et al. 2010). The rate of encounter will vary with the size of the river plume, principally how far river water extends away from the coast and how far from shore sea lamprey traverse the coast in that region. The size of river plumes vary with river discharge, coastal bathymetry, coastline configuration at the river mouth (e.g., embayment trapping water), and wind driven lake current (Churchill et al. 2003, Rao and Schwab 2007). Wind-driven current and the coastal boundary are the dominant features regulating the extent of the river plume in the Great Lakes (Churchill et al. 2003). River plume size is likely a key factor but is not explicitly linked with river discharge a potential explanation to why Moore and Schleen (1980) failed to see a relationship between sea lamprey entry and river discharge for rivers that were known to contain sea lamprey larvae. Sea lamprey traversing the coast at 300 m from shore in search of river water may miss many smaller river plumes, such as the Black Mallard River plume that was observed extending less than 250 m from shore on 80 % of nights in spring (Meckley et al. 2014b). The number of sea lamprey available to encounter the plume will depend on the number of sea lamprey to return to the particular region after the parasitic stage and whether they reach a given river or stop at a previous river. It remains unclear if the start of the migration could bias encounter with a specific region of a lake or if release from a host occurs broadly throughout the lake without major patterns, as virtually nothing is known about the behavior of parasitic sea lamprey beyond wounding rates (Jorgensen and Kitchell 2005, Silva et al. 2014). Additional research evaluating the conditions associated with the number of sea lamprey to enter a river would be valuable. 183 Movement in the river plume Once in the river plume sea lamprey regularly moved towards and away from the coast as they made persistent progress along the coast; they reversed direction in shallow water and at consistent distances from the coast, a pattern termed coastal rebounding (Meckley et al. 2014) (Figure 4.11, Figure 4.12). We hypothesize that coastal rebounding relies on thigmotaxis within an odor defined region of the coast, where sea lamprey turn once reaching the edges of the river plume and the physical coastal edge to progress up and down the coastline. The behavior is consistent with localization and assessment, as sea lamprey often approached very close to the river mouth on multiple occasions without entering the river. A river mouth will almost always occur near one end of the river plume although some lake conditions can consistently create a more centrally located river mouth with respect to the river plume (García Berdeal et al. 2002, Churchill et al. 2003), and the river mouth may not be located at the strongest concentration of river water in cases where water is trapped within a basin. Coastal rebounding would allow for localization of the river mouth regardless of the dynamic river plume conditions. Sea lamprey do not follow the coast in shallow water but instead repeatedly move toward the coast and offshore as progress is made along the coast. There are at least three explanations. First the rebounding nature may enable sea lamprey to avoid entrapment in complex shorelines but second, it could result in avoidance of habituation to olfactory information in the river plume. Sensory habituation can occur if a receptor is constantly exposed to a stimulus and becomes fatigued resulting in a loss of sensitivity (Ferrari et al. 2010). Olfactory habituation can be induced by multiple paradigms, although the type of habituation this type of movement would best avoid is habituation by a change in the function of the olfactory bulb, a process which takes longer than 30 minutes (Chaudhury et al. 2010). For example sea lamprey habituate to the odor 184 of dead adult sea lamprey after exposure for between two and four hours (Jason Bals, unpublished data). To address whether sea lamprey are reaching the edge of the river plume, a more accurate representation of the river plume than a dye concentration model would be required, which could be provided by a particle diffusion model as a large number of particles provide a more reliable edge structure (Mantha S. Phanikumar, personal communication). Third, the search could be a sampling period and sea lamprey could be collecting other information such as the presence of mates. Given that sea lampreys conceptually need to locate two things, spawning habitat and mates; it seems that recognition of mates could be important. The role of larval odor Our leading hypothesis is that once encountering a river plume individuals either enter, stage, or search for a different river based on the quality of the river and river temperature (Binder et al. 2010, Clemens et al. 2010, Vrieze et al. 2011, Meckley et al. 2014b). Odor and in particular larval odor has support from lab and field tests as an important cue that informs a migrant of the quality of a river (Moore and Schleen 1980, Vrieze and Sorensen 2001, Vrieze et al. 2010, Meckley et al. 2014b). In this study sea lamprey entered the Ocqueoc River at a higher rate in 2011 than 2010, two years removed from any TFM treatment. We saw no evidence that an environmental variable was associated with the way sea lamprey moved in the river plume or entry rate, although features like river discharge were higher in 2011 (discharge mean (m3·S-1) (minimum, maximum) 2010: 2.64 (0.6-9.7); 2011: 3.9 (1.8-8.6)) and river temperature varied between years (Figure 4.13). Our observations of reduced entry after treatment is consistent with historical observation in the Ocqueoc River (Figure 4.14), and is a wide spread pattern throughout rivers in the Great Lakes (Moore and Schleen 1980). These observations provide 185 additional support that river water alone is sufficient to induce local search, while larval odor influences the river entry decision. Search behavior did not differ between years, consistent with the hypothesis that river water alone induced the transition from extensive search to local search in the river plume and larval odor influenced the river entry decision. This could only have been confirmed if no sea lamprey larvae were in the river in 2010, which was not the case in this study (Personal communication Aaron Jubar of the U.S Fish and Wildlife Service). Previous suggestions that sea lamprey moved faster outside of the river plume, searched the river plume through klinokinesis (Vrieze et al. 2011), and stopped more frequently in the river plume (Meckley et al. 2014b), were unsupported and likely the result of inaccuracies in manual acoustic telemetry and short-term observations. The interpretation that sea lamprey moved slower in the river plume was likely a result of the order of observations as activity in the river plume was often captured at the start of the night and movement outside of the river plume was observed later in the night. This resulted in the slow ramp up in speed we observed after settlement being attributed to movement in the river plume. The in plume search does not appear to be a kinesis as there is clear persistence and reliance on the coastline during search. The persistent exploration up and down the coastline in the river plume is consistent with an absolute olfactory search mechanism where the presence of river water defines the search boundary, rather than a differential mechanism where the cue intensity is important and gradients are used to provide directional information (Benhamou and Bovet 1989). The pattern of movement is inconsistent with reliance on odor gradients for navigation but we are unable to fully rule out chemotaxis as increases in a chemical cue gradient while rebounding could inform the migrant if it is moving towards a river mouth and could be undetected in our study as sea lamprey do make repeated passes up and down the coast. For 186 example, Blue crabs (Callinectes sapidus) use a combined mechanism in which they locate odor sources with rheotaxis but require chemotaxis to adjust to gradients across the current stream (Zimmer-Faust et al. 1995). River entry The decision of a sea lamprey to enter a river could occur prior to reaching the river mouth, though in the Ocqueoc River when other migrants were present and spawning and rearing habitat are available, individuals were never observed reversing course after moving past the immediate river mouth. Sea lamprey that reached the immediate river mouth did not always enter. In both years sea lamprey came to the immediate river mouth without entering and encountered river water on multiple nights, though in 2011 sea lamprey that did either task entered at a high rate (Figure 4.6). River entry most often occurred following a series of very tight rebounds along the coast before moving into the river mouth. Sea lamprey transition from rebounding movement to straight upstream movements on the bottom, suggesting that sea lamprey transition from edge guided search to odor mediated rheotaxis as they enter the river (Vrieze et al. 2011, Johnson et al. 2012, Meckley et al. 2014b). Vertical excursions also cease at this point in the migration (Chris Holbrook, personal communication). Two previously reported aspects of entry were not regularly observed in this study, including staging in front of the Ocqueoc River or repeated entries and exits at the river mouth (Applegate 1950). Sea lamprey rarely remained stationary during nautical twilight and only in three instances of 151, they entered the river mouth and then exited. One of the three individuals reentered the river. This does not support or discredit the occurrence of staging or repeated entry as the sea lamprey in this study had already entered a river and the river temperatures were between 10-20 C ̊ for the majority of the study period, well above the 4.4 C ̊ temperature in 187 which these behaviors were reported by Applegate (1950) and above the 10 C ̊ thermal threshold for active migration in the river (Binder and McDonald 2008). Given that many individuals were reported staging together, and sea lamprey often spent multiple days moving in the plume in front of the river, there is a possibility that social information is important to the decision to enter a river. Sea lamprey would not have to actively follow a sea lamprey, by maintaining proximity to an individual, to use social information to enter a river. Social information could come in the form of visual observations of other individuals or odor recognition of mates in the area. The odor or visual recognition of mates could be an important piece of information, as it is known that the odor of maturing migrants synchronizes the migration (Chung-Davidson et al. 2013). Although channel bias was not observed in immature female lamprey in response to the male released sex pheromone (Siefkes et al. 2005), migrants are believed to increase upstream migration in the odor’s presence (Johnson et al. 2013), suggesting that odor-based mate recognition is possible. Odor mediated manipulation of sea lamprey habitat selection In addition to the goal of manipulation of sea lamprey river selection by encouraging sea lamprey to enter streams with poor spawning or rearing habitat as suggested by Li et al. (2007), we suggest that a second potential method would be encouraging sea lamprey to enter a river the year after TFM application, in rivers treated every three years. Ideally recruitment could be raised in the year after treatment to a rate similar to the second or third year entry rates following treatment. It is unclear if the high rate of recruitment in the third year is due to a further increase in entry between the second and third year post TFM treatment or if additional encounter occurs due to TFM treatment of other neighboring rivers on a different pesticide treatment schedule. The only weak comparison for evaluating the effect of the three synthetic compounds found in 188 larval odor provided no evidence of an effect on river entry in this study. Combined with other field tests it appears that the full larval odor or at minimum additional or different components found in larval odor will be required to manipulate tributary selection or river entry (Meckley et al. 2012, Meckley et al. 2014b). Summary The sea lamprey lake migration has now begun to be characterized over a narrow region of Lake Huron in Hammond Bay (Vrieze et al. 2010, Vrieze et al. 2011, Meckley et al. 2014b), suggesting the evaluation of the migration on other coastlines could be a valuable future contribution. All paths of sea lamprey movement observed on the first night that did not encounter river water and did encounter river water can be observed in Figure 4.15 and Figure 4.16, respectively. This study represents the finest scale exploration of in plume search to date, however an improvement would be the development of a hydrodynamic model that depicts the plume edges more finely and captures the exact role of the river plume edge in the search process. Notably, do sea lampreys exit the plume before counterturning back toward the coast or do sea lamprey reverse back towards the coast once reaching a physical feature of the water column? Observing the behavior of sea lamprey near the end of the river plume not near the river mouth could also be informative as our array only offered a window into the river mouth side of plume search. It will be difficult to reveal a chemotactic search mechanism in the river plume even if it exists because sea lamprey often make many passes up and down the coast, providing the impression that the movement is not explicitly in the river plume or performed to localize the river mouth but may be a direct assessment of the river quality (e.g., searching for larval odor or presence of mates). Sea lamprey appear to be a physical edge-guided species, that relies heavily on the surface, bottom, and coastal edges to guide the migration first to the coast 189 (Chapter 3), and then along the coast in search of rivers. Simple movement rules leading sea lamprey to the coast, along the coast, and within a river plume, appear sufficient to guide sea lamprey to a river mouth in a similar way that simple movement rules guide other species to geographically distinct spawning habitat and to find mates (e.g. Pe'er et al. 2004). These findings suggest that the currently identified components of larval odor are insufficient to influence river entry, although a more thorough test would be completed in a river lacking any larval odor of any lamprey species. Given the likely lack of coevolution that would bias how sea lamprey larvae release the larval signal, it is possible that components of larval odor that are used by migrants is composed of a more complex number of components than would be common to a sex pheromone and may be harder to replicate. However our observations do suggest that the use of the entire synthetic larval odor could be a powerful management tool, a designation that appears less likely for the male released sex pheromone. ACKNOWLEDGEMENTS This research was funded through grants from the Great Lakes Fishery Commission including by way of the Great Lakes Restoration Initiative appropriations (GL-00E23010-3) and was accomplished through considerable assistance from the US Fish and Wildlife Service and the USGS Hammond Bay Biological Station. 190 APPENDIX 191 Table 4.1: Three specific criteria were adopted to establish the data quality objectives based on an extensive review of data quality (Meckley et al. 2014a). 1 2 3 Objective Criteria Detect changes in Mean error ≤1.77 trajectory m Rationale Many trajectory based analyses are only preserved when the average position error is <10% of the mean step length (Bradshaw et al. 2007). Transmission delay was (33.4 s), sea lamprey ground speed is 0.53 ms−1 (Vrieze et al. 2011, Meckley et al. 2014b), equating to a step length of 17.7 m. Assign behavioral Max error <15 m Simple behavioral assignment of moving or state (moving or stationary with first passage time analysis, which is a stationary) measure of the time it takes an individual to leave a circle of fixed radius r drawn around each measured location to determine if movement is occurring (Fauchald and Tveraa 2003, Barraquand and Benhamou 2008). Balance loss of Retain 95% of To ensure representation of fish tracks while acceptable data acceptable data, avoiding loss of accurate data. We aimed to keep and retention of Reject ≥99% of 95% of acceptable data (≤6 m) while retaining <1% unacceptable data unacceptable data of unacceptable data (>6 m) of all retained positions. 192 Table 4.2: Criteria for selection of an HPE filter cutoff. In the data values listed for each criterion the HPE cutoffs are listed acceptable HPE cutoffs are listed for 2010 above and 2011 below. Criteria Stationary Mobile Rational for selection of a HPE filter cutoff of 8 test test 1 2 All All 3 This criterion was met for stationary tests. The mobile test Criterion 1 Mean ≤1.77 m All NA None would have required a very low HPE to attain below 1.77 m. Although for a number of reasons stationary tests are more reliable estimators for this metric (Meckley et al. 2014a). <7 <8 ≤19 In 2010: An HPE of 8 did not meet the criteria for mobile Criterion 2 tests, but only 1 of 134 positions remained. The stationary Max error <15 m <7 NA ≤1 test criteria were met. Only 2 of 45,744 positions were problematic (<0.001%) for the combined stationary tests (Figure 4.4). In 2011 an HPE of 7 removed all except one position for the stationary test, and one position for the mobile test exceeding 15 m error. 8–10 3–15 None In 2010: An HPE cutoff of 8 met the criteria for both Criterion 3 % incorrectly 3-15 NA None stationary tests, and although the mobile test did not have a retained vs. % suitable range, the range from 5–10 was equally effective. In incorrectly 2011: An HPE of 7 met criteria for the stationary test as it rejected fell within at least 3-15. 193 Table 4.3: Akaike’s information criterion (AIC) values for linear mixed effects models using river entry (Em) as a binary response variable to various combinations of fixed effects including year (Y), iterative number of nights the sea lamprey encountered river water (N), a year by night number interaction (YxN), average nightly discharge of the Ocqueoc River (DC: m3·S-1), average lake temperature encountered (LT: C ̊), average nightly river temperature (RT; C ̊ ), the difference in experienced lake temperature and river temperature (DFT; C ̊ ), minimum distanced reached from the river mouth (D: km), and average water velocity experienced in the lake (V). Model Fixed Effects df Vm AIC (ΔAIC) 0 1 2 209.8 (13.0) 1 Y 3 202.6 (5.8) 2 N 3 211.7 (14.9) 3 Y + YxN 5 205.7 (8.9) 4 DC 3 211.8 (15.0) 5 Y + DC 4 204.5 (7.7) 6 LT + RT 4 207.5 (10.7) 7 Y + LT 4 199.7 (2.9) 8 Y + RT 4 204.3 (7.5) 9 Y + DFT 4 199.6 (2.8) 10 Y + LT + D 5 197.1 (0.3) 11 Y + DFT +D 5 196.8 (0) 12 Y + DFT +LT 5 201.5 (4.7) 13 14 Y + DC + LT + RT Y+V 6 6 203.5 (6.7) 199.2 (2.4) 15 Y + LT + RT + V +DC + D + YxN 10 205.8 (9) 194 Table 4.4: Akaike’s information criterion (AIC) values for a nonlinear mixed-effects model in the formulation described by Lindstrom and Bates (1990), using the function “nmle” in R (R Development Core Team 2015). Ground speed (Vm) as a response variable to various combinations of fixed effects including year (Y), the log of the cumulative distance traveled (D), whether encounter with river water occurred yet in the path (E), or if the point was currently in the plume (P), temperature at fish position (T), and an interaction effect between water current magnitude and direction by fish heading interaction (CxF); and finally the random effect of individuals (animal ID). Model Fixed Effects Df Vm AIC (ΔAIC) 0 E 5 178.3 (99.2) 1 E + E:D 6 89.2 (10.1) 2 Y + E:D 5 276.2 (197.1) 3 T 5 279.5 (200.4) 4 P +Y 6 290.9 (211.7) 5 T+P+E 7 168.9 (89.4) 6 T+P+Y 7 279.7 (200.6) 7 T + CxF 6 278.1 (199.0) 8 P + CxF + E:D 8 98 (19.7) 9 T + P + CxF + E + E:D 9 79.1 (0) 195 Table 4.5: Akaike’s information criterion (AIC) values for a nonlinear mixed-effects model in the formulation described by Lindstrom and Bates (1990), using the function “nmle” in R (R Development Core Team 2015). Path straightness (Sm) as a response variable to various combinations of fixed effects including year (Y), the log of the cumulative distance traveled (D), if whether encounter with river water occurred yet in the path (E), or the point was currently in the plume (P); and finally the random effect of individuals (animal ID). Model 6 and 9 had the lowest AIC scores. Model Fixed Effects Df Vm AIC (ΔAIC) 0 D+Y 6 -354.6 (7.5) 1 D+E 6 -358.0 (4.1) 2 E+Y 7 -340.4 (21.7) 3 P+Y 7 -348.9 (13.2) 4 P+E 6 -355.2 (6.9) 5 E 5 -353.2 (8.8) 6 P 5 -360.4 (1.7) 7 D+E +P 7 -356.8 (5.3) 8 D+E+P+Y 8 -351.9 (10.2) 9 DxE + P -362.0 (0) 196 Table 4.6: To evaluate the effect of a host of variables on river entry (Em), a mixed effects logistic regression with a binary response variable (river entry: 1 or 0) was implemented in R using glmer (lme4 package, Bates et al. 2014). The best fitting models, as determined by AIC, R code: m11<-glmer(Em ~ Y+ DFT+ D+ (1|ID), data=entry, binomial) Family: binomial ( logit ) Formula: Em ~ Y + DFT + D + (1 | ID) Data: entry AIC 196.8 BIC logLik deviance df.resid 211.9 -93.4 186.8 144 Scaled residuals: Min 1Q Median 3Q Max -1.1786 -0.7532 -0.5324 0.9475 2.8220 Random effects: Groups Name ID (Intercept) Number of obs: Variance Std.Dev. 1e-14 1e-07 149, groups: ID, 98 Fixed effects: (Intercept) year2011 DFT D --Signif. codes: Estimate -0.64956 0.88611 0.02464 -2.08616 Std. Error 0.41248 0.35433 0.04996 1.05819 z value -1.575 2.501 0.493 -1.971 0 ‘***’ 0.001 ‘**’ 0.01 ‘*’ 0.05 ‘.’ 0.1 ‘ ’ 1 Correlation of Fixed Effects: (Intr) yr2011 year2011 -0.526 DFT -0.615 0.017 M -0.436 0.050 DFT -0.015 197 Pr(>|z|) 0.1153 0.0124 * 0.6219 0.0487 * Table 4.6: (cont’d) for river entry included year and minimum distance reached from the river and either the lake temperature experienced (Model 10) or the difference in lake temperature experienced and the river temperature (Model 11). The fixed effects included year (Y), iterative number of nights encountering river water (N), a year by night number interaction(YxN), average nightly discharge of the Ocqueoc River (DC: m3·S-1), average lake temperature encountered (LT: C ̊), average nightly river temperature (RT; C ̊ ), the difference in experienced lake temperature and river temperature(DFT; C ̊ ), minimum distanced reached from the river mouth (D: km), and average water velocity experienced in the lake (V). Here is Model 11 output. 198 Table 4.7: Output from the summary of the best fit nonlinear mixed-effects model for the continuous response variable ground speed (body lengths per second), Model 9. The model Linear mixed-effects model fit by REML Data: GirardEnc AIC BIC logLik 79.13117 146.8898 -29.56558 Random effects: Formula: ~1 | ID (Intercept) Residual StdDev: 0.2078676 0.2983834 Correlation Structure: AR(1) Formula: ~1 | ID Parameter estimate(s): Phi: 0.5945631 Fixed effects: Vm ~ T+ P + C:F + E + I(D/1000): E Value Std.Error DF t-value (Intercept) 0.6803459 0.04771404 6437 14.258818 t 0.0198588 0.00355488 6437 5.586343 as.factor(P)Y 0.0018479 0.01246258 6437 0.148279 as.factor(E)b 0.3591986 0.02492367 6437 14.411945 C:F -0.2506198 0.11486012 6437 -2.181956 Ea:I(D) 0.1645548 0.01549279 6437 10.621383 Eb:I(D) -0.0016136 0.00364662 6437 -0.442490 Correlation: t PY E C:F Ea:D Eb:D (Intr) -0.606 0.167 -0.297 -0.006 -0.238 0.021 t -0.250 0.040 -0.163 -0.021 -0.046 P -0.273 -0.027 0.007 0.107 Standardized Within-Group Residuals: Min Q1 Med Q3 Max -4.9337612 -0.4086086 0.1300112 0.6036556 8.6886942 Number of Observations: 6483 Number of Groups: 40 199 p-value 0.0000 0.0000 0.8821 0.0000 0.0291 0.0000 0.6581 E C:F Ea:D 0.042 0.458 -0.482 0.042 -0.017 0.073 Table 4.7: (cont’d) included the temperature experienced (T), whether encounter with the river plume had occurred in the path (E), whether the sea lamprey was in the river plume (P), and an interaction between water current and if the fish was moving towards or away from the current direction. The correlation statement accounts for first order autocorrelation in the residuals that occurred from analyzing multiple simultaneous steps relating to the same individual. The random effect of animal ID is included. 200 Table 4.8: Output from the summary of the best fit nonlinear mixed-effects model for path straightness (0-1), Model 9. The model included the effect of the interaction between cumulative R code: m9 <- lme(fixed = S ~ I(D/1000):as.factor(E) + as.factor(P), random = ~ 1 | ID, correlation = corAR1(form = ~ 1|ID), data = sin) Linear mixed-effects model fit by REML Data: sin AIC BIC logLik -362.0493 -329.653 188.0247 Random effects: Formula: ~1 | ID (Intercept) StdDev: 0.06001644 Residual 0.1812923 Correlation Structure: AR(1) Formula: ~1 | ID Parameter estimate(s): Phi 0.1026954 Fixed effects: sin ~ I(D/1000):as.factor(E) + as.factor(P) Value Std.Error DF (Intercept) 0.7951321 0.01479099 684 as.factor(P)1 -0.0450035 0.01790223 684 I(D/1000):as.factor(E) 0 0.0482397 0.01201762 684 I(D/1000):as.factor(E)1 0.0179331 0.00434604 684 Correlation: as.factor(plume)1 I(stepcum/1000):as.factor(enc) I(stepcum/1000):as.factor(enc)1 (Intr) -0.318 0 -0.415 t-value 53.75785 -2.51385 4.01408 4.12631 p-value 0.0000 0.0122 0.0001 0.0000 as.()1 I(/1000):.()0 -0.552 -0.262 Standardized Within-Group Residuals: Min Q1 Med Q3 Max -4.1849828 -0.3403049 0.3338883 0.6488679 1.5624139 Number of Observations: 760 Number of Groups: 73 201 0.206 0.324 Table 4.8: (cont’d) distance traveled (D) and Encounter with the river plume and the fixed effect of being in the river plume. The correlation statement accounts for first order autocorrelation in the residuals that occurred from analyzing multiple simultaneous steps relating to the same individual. The random effect of animal ID is included. The distance is divided by 1000 to put the variable in km rather than m for easier interpretation. 202 Figure 4.1: The study site was located in Hammond Bay of Lake Huron (A, B). Sea lamprey were released at three locations (open circles). The study site was covered by a two kilometer wide array that reached a kilometer from shore (C). The receiver positions in 2010 are represented by triangles, and in 2011 they are represented by circles. In each year nine sync tags were distributed in two rows across the array. During peak performance a transmitter could be positioned every 15-45 s. Two of the six river receivers used for monitoring passage are represented by black squares. The receivers in the river could not detect tags in the lake. 203 Figure 4.2: The hydrodynamic model makes predictions of the current speed and direction and lake temperature at each node (N=4346) of an unstructured grid fit to Hammond Bay (FVCOM grid).The size of the triangular grid cells decreases near the river mouth to provide a higher precision near our fish tracking array (minimum grid size: 9.15 m2). The hydrodynamic model provides predictions of current speed and direction and water temperature. The model is three dimensional and provides predictions in 2010 and 2011 at 20 slices through the water column water column, independent of depth (m), which is color coded. 204 Figure 4.3: Comparisons between vertical averaged currents form ADCP observations and simulations taken in 2011. 205 Figure 4.4: Comparisons between observed temperature recorded by the ADCP and hydrodynamic model simulations in 2011. 206 Figure 4.5: Sea lamprey mostly began moving between sunset (orange vertical lines) and nautical twilight NT (gray box) in most cases (A). Activity continued until between NT and sunrise, although activity continued in lighter conditions than under which it commenced (B). The double vertical lines represents sunrise and sunset as the time from NT changes with night length. Sea lamprey entered the river throughout the night but the earliest they were observed entering was just before NT (C) and despite activity being observed at sunrise in the lake, river entry was never observed outside of NT in the lake (D). Two graphs show river entry as night length changes with day length, this explains the dark gray areas (NT, all year), and light gray areas (NT during some parts of the year). Focus on the 0 side of insets C and D. 207 ≥ Figure 4.6: The entry rate based on the total number of different nights that sea lamprey were observed encountering river water in 2010 and 2011. The nights do not have to be consecutive and don’t have to include the first night after release. The entry rate on the first night “1st night” refers to the entry rate of sea lamprey that encountered river water on the first night of release after settlement. The number of sea lamprey that fall into each category for each year are found at the top of the graph. 208 Northing Easting Figure 4.7: The 33 tracks of sea lamprey that did not encounter river water on the first night are color coded and the arrow of the heading from the 10th to last point to final point are shown (A). 209 Figure 4.7: (cont’d) The circular plot in A (top) shows that most fish moved west out of the array. The exit arrow (thick arrow) is shown for each sea lamprey and the average current experienced is shown by the thin arrow, where the arrow points the current direction and the length represents current strength (max: 27 cm·s-1,average: 4.9 cm·s-1) (B). The circle plots show the current directions observed on the 33 nights (“Current”), the direction sea lamprey moved when the current was from the East (“Fish-East Current”), the direction the sea lamprey moved when the current was from the west (“Fish-West Current”). 210 Figure 4.8: Each transect (set of two points) for sea lamprey that never encountered river water were binned by 100 m increments from the coast based on their approach ratio (distance moved towards the coast/total distance traversed). When sea lamprey were 500-1000 m from the coast movement was directed towards the coast (1) before turning and moving parallel to the coast (0500 m). An increase in movement away from the coast occurs between 100-300 m from the coast resulting in an average parallel movement along the coast of 303 m ± 221 m when exiting the array (centered at an approach ratio of 0). 211 Figure 4.9: Ground speed (Body lengths per second) is shown for each step over the cumulative path length for all sea lamprey (N=33) that did not encounter river water on the first night. Each line represents a different individual. The pattern for before river plume encounter (black circles) and after river plume encounter (red squares) is shown. Observations are restricted to a cumulative path length of 2500 meters. A ramp up in speed to one body length per second is evident in the first 500 meters of activity. 212 Figure 4.10: The google earth image depicts the distinct nearshore wave zone which includes a comparatively steep series of banks and troughs that run horizontal to the coast compared the shallow bathymetric gradient in Hammond Bay. The image captures the coastline in Hammond Bay in front of the Ocquoec River Mouth, located in the bottom right of the image. The edge of the banked zone extends from 204-308 m from shore in this area of the array. 213 Figure 4.11: A smoothed utilization distrution in 50 m cells of all active sea lamrpey points following river plume encounter as defined by encounter with a dye concentraiton of greater than 5 ppm in 2010 and 2011. The white line represents the Ocquoec River. Movement extended farther east and west along the coast in both cases but was not captured by the array. 214 Figure 4.12: The river plume occupied three general forms in Hammond Bay as represented by the output of the dye concentration model. The image to the left represents a less common river plume extending to the west. The central schematic depicts the river plume that occurs when the plume switches from either east to west or occasionally when currents were moving in northerly or southerly directions and pushed the plume out from shore. The most common river plume extended to the east and extended out of Hammond bay. The river water was often trapped to the east and west due to the shape of the coast. 215 Figure 4.13: The temperature by Julian date for 2010 (Blue Line) and 2011 (Orange Line) is shown. A symbol represents each fish if it was in the array on a given night in 2010 (Blue) and 2011 (Orange) at the maximum temperature encountered by the fish. Triangles represent fish that entered the river on that night and circles representing fish that do not enter that night. An individual fish is represented multiple times if it returns on multiple nights. 216 Figure 4.14: The percentage of the entire population of adult sea lamprey in Lake Huron that return to the Ocqueoc River to spawn from 1995 to 2015 based on mark recapture (black circles) and the change in entry between years (blue circles). A change of entry rate of 2 represents a doubling in the rate of entry. Vertical lines represent the times of TFM treatment to the Ocquoec River. The change in entry rate from one year after TFM treatment to the second year shows that the increase in entry rate between 2010 and 2011 (D) fits within the historical increase in rate (A: 1.8, B: 1.3, C: 3.1, D: 2.1). 217 Time Figure 4.15: The 40 sea lamprey that passed through the river plume based on the estimates of the dye concentration at each fish position are shown in a schematic of Hammond bay, where the black line represents the coast and the red dot represents the river mouth. Each page is labeled by the year (10 or 11) and animal ID. The upper graph shows the sea lamprey movement on the first night from where the sea lamprey settled in the array after release (orange dot). In the schematic of movement, the gray arrow points the heading during the last 5 points of the track and the track 218 Figure 4.15: (cont’d) is color coded by 15 minute intervals of activity. The label of “Entered” indicates that sea lamprey eventually entered the Ocqueoc River and the label of “Did Not Enter” indicates that the sea lamprey never entered the Ocqueoc River. The number of bouts follows the label indicating the number of different nights the sea lamprey was observed in the array. If the sea lamprey Time 219 Figure 4.15: (cont’d) entered and the bout is one, the individual entered at the end of the track shown. In some cases entry is not captured well by the array (e.g. 10-17). The lower graph depicts the depth of the sea lamprey during movement through time and the color of the points corresponds to the schematic in the upper graph. The red line represents the estimated total water column depth. When the fish is below the line it can be assumed the fish is on the bottom. Time 220 Figure 4.15: (cont’d) Time 221 Figure 4.15: (cont’d) Time 222 Figure 4.15: (cont’d) Time 223 Figure 4.15: (cont’d) Time 224 Figure 4.15: (cont’d) Time 225 Figure 4.15: (cont’d) Time 226 Figure 4.15: (cont’d) Time 227 Figure 4.15: (cont’d) Time 228 Figure 4.15: (cont’d) Time 229 Figure 4.15: (cont’d) Time 230 Figure 4.15: (cont’d) Time 231 Figure 4.15: (cont’d) Time 232 Figure 4.15: (cont’d) Time 233 Figure 4.15: (cont’d) Time 234 Figure 4.15: (cont’d) Time 235 Figure 4.15: (cont’d) Time 236 Figure 4.15: (cont’d) Time 237 Figure 4.15: (cont’d) Time 238 Figure 4.15: (cont’d) Time 239 Figure 4.15: (cont’d) Time 240 Figure 4.15: (cont’d) Time 241 Figure 4.15: (cont’d) Time 242 Figure 4.15: (cont’d) Time 243 Figure 4.15: (cont’d) Time 244 Figure 4.15: (cont’d) Time 245 Figure 4.15: (cont’d) Time 246 Figure 4.15: (cont’d) Time 247 Figure 4.15: (cont’d) Time 248 Figure 4.15: (cont’d) Time 249 Figure 4.15: (cont’d) Time 250 Figure 4.15: (cont’d) Time 251 Figure 4.15: (cont’d) Time 252 Figure 4.15: (cont’d) Time 253 Figure 4.15: (cont’d) Time 254 Figure 4.15: (cont’d) Time 255 Figure 4.15: (cont’d) 256 Figure 4.15: (cont’d) Time 257 Time Figure 4.16: The 33 sea lamprey that did not encounter the river plume on the first night based on the estimates of the dye concentration at each fish position are shown in a schematic of Hammond bay, where the black line represents the coast and the red dot represents the river mouth. Each page is labeled by the year (10 or 11) and animal ID. The upper graph shows the sea lamprey movement on the first night from where the sea lamprey settled in the array after release (orange dot). In the schematic of movement, the gray arrow points the heading during the 258 Figure 4.16: (cont’d) last 5 points of the track and the track is color coded by 15 minute intervals of activity. The label of “Entered” indicates that sea lamprey eventually entered the Ocqueoc River and the label of “Did Not Enter” indicates that the sea lamprey never entered the Ocqueoc River. The number of bouts follows the label indicating the number of different nights the sea lamprey was observed in Time 259 Figure 4.16: (cont’d) the array. If the sea lamprey entered and the bout is one, the individual entered at the end of the track shown. In some cases entry is not captured well by the array (e.g. 10-17). The lower graph depicts the depth of the sea lamprey during movement through time and the color of the points corresponds to the schematic in the upper graph. The red line represents the estimated total water column depth. When the fish is below the line it can be assumed the fish is on the bottom. Time 260 Figure 4.16: (cont’d) Time 261 Figure 4.16: (cont’d) Time 262 Figure 4.16: (cont’d) Time 263 Figure 4.16: (cont’d) 264 Figure 4.16: (cont’d) 265 Figure 4.16: (cont’d) Time 266 Figure 4.16: (cont’d) Time 267 Figure 4.16: (cont’d) Time 268 Figure 4.16: (cont’d) Time 269 Figure 4.16: (cont’d) Time 270 Figure 4.16: (cont’d) Time 271 Figure 4.16: (cont’d) Time 272 Figure 4.16: (cont’d) Time 273 Figure 4.16: (cont’d) Time 274 Figure 4.16: (cont’d) Time 275 Figure 4.16: (cont’d) Time 276 Figure 4.16: (cont’d) Time 277 Figure 4.16: (cont’d) Time 278 Figure 4.16: (cont’d) Time 279 Figure 4.16: (cont’d) Time 280 Figure 4.16: (cont’d) Time 281 Figure 4.16: (cont’d) Time 282 Figure 4.16: (cont’d) Time 283 Figure 4.16: (cont’d) Time 284 Figure 4.16: (cont’d) Time 285 Figure 4.16: (cont’d) Time 286 Figure 4.16: (cont’d) Time 287 Figure 4.16: (cont’d) Time 288 Figure 4.16: (cont’d) Time 289 Figure 4.16: (cont’d) Time 290 REFERENCES 291 REFERENCES Able, K. P. (1991). "Common Themes and Variations in Animal Orientation Systems." American Zoologist 31(1): 157-167. Applegate, V. C. (1950). The natural history of the sea lamprey in Michigan. Washington, D. C, U. S. Department of Interior Fish & Wildlife Service, Special Scientific Report Fisheries. Barraquand, F. and S. Benhamou (2008). "Animal movements in heterogeneous landscapes: identifying profitable places and homogenous movement bouts." Ecology 89: 3336 - 3348. Bates, D., M. Maechler, B. M. Bolker and S. Walker (2014). lme4: Linear mixed-effects models using Eigen and S4. R package version 1.1-7. Beletsky, D., J. H. Saylor and D. J. Schwab (1999). "Mean Circulation in the Great Lakes." Journal of Great Lakes Research 25(1): 78-93. Benhamou, S. and P. Bovet (1989). "How animals use their environment: a new look at kinesis." Animal Behaviour 38(3): 375-383. Bergstedt, R. A. and J. G. Seelye (1995). "Evidence for Lack of Homing by Sea Lampreys." Transactions of the American Fisheries Society 124(2): 235-239. Binder, T. R. and D. G. McDonald (2007). "Is there a role for vision in the behaviour of sea lampreys (Petromyzon marinus) during their upstream spawning migration?" Canadian Journal of Fisheries and Aquatic Sciences 64(10): 1403-1412. Binder, T. R. and D. G. McDonald (2008). "The role of temperature in controlling diel activity in upstream migrant sea lampreys (Petromyzon marinus)." Canadian Journal of Fisheries and Aquatic Sciences 65(6): 1113-1121. Binder, T. R., R. L. McLaughlin and D. G. McDonald (2010). "Relative Importance of Water Temperature, Water Level, and Lunar Cycle to Migratory Activity in Spawning-Phase Sea Lampreys in Lake Ontario." Transactions of the American Fisheries Society 139(3): 700-712. Bjerselius, R., W. Li, J. H. Teeter, J. G. Seelye, P. B. Johnson, P. J. Maniak, G. C. Grant, C. N. Polkinghorne and P. W. Sorenson (2000). "Direct behavioral evidence that unique bile acids 292 released by larval sea lamprey (Petromyzon marinus) function as a migratory pheromone." Canadian Journal of Fisheries and Aquatic Sciences 57: 557-569. Bradshaw, C., D. Sims and G. Hays (2007). "Measurement error causes scale-dependent threshold erosion of biological signals in animal movement data." Ecol Appl 17(2): 628 - 638. Chaudhury, D., L. Manella, A. Arellanos, O. Escanilla, T. A. Cleland and C. Linster (2010). "Olfactory bulb habituation to odor stimuli." Behavioral neuroscience 124(4): 490-499. Chung-Davidson, Y. W., H. Y. Wang, M. J. Siefkes, M. B. Bryan, H. Wu, N. S. Johnson and W. M. Li (2013). "Pheromonal bile acid 3-ketopetromyzonol sulfate primes the neuroendocrine system in sea lamprey." Bmc Neuroscience 14: 13. Churchill, J. H., E. A. Ralph, A. M. Cates, J. W. Budd and N. R. Urban (2003). "Observations of a negatively buoyant river plume in a large lake." Limnology and Oceanography 48(2): 884-894. Clemens, B. J., T. R. Binder, M. F. Docker, M. L. Moser and S. A. Sower (2010). "Similarities, Differences, and Unknowns in Biology and Management of Three Parasitic Lampreys of North America." Fisheries 35(12): 580-594. Dodge, K. L., B. Galuardi, T. J. Miller and M. E. Lutcavage (2014). "Leatherback Turtle Movements, Dive Behavior, and Habitat Characteristics in Ecoregions of the Northwest Atlantic Ocean." PLoS ONE 9(3): e91726. Fauchald, P. and T. Tveraa (2003). "Using first-passage time in the analysis of area-restricted search and habitat selection." Ecology 84(2): 282 - 288. Fine, J. M. and P. W. Sorensen (2008). "Isolation and biological activity of the multi-component sea lamprey migratory pheromone." Journal of chemical ecology 34(10): 1259-1267. Galperin, B., L. H. Kantha, S. Hassid and A. Rosati (1988). "A Quasi-equilibrium Turbulent Energy Model for Geophysical Flows." Journal of the Atmospheric Sciences 45(1): 55-62. García Berdeal, I., B. M. Hickey and M. Kawase (2002). "Influence of wind stress and ambient flow on a high discharge river plume." Journal of Geophysical Research: Oceans 107(C9): 3130. 293 Gore, J. A. (2007). Chapter 3 - Discharge Measurements and Streamflow Analysis. Methods in Stream Ecology (Second Edition). F. R. Hauer and G. A. Lamberti. San Diego, Academic Press: 51-77. Gurarie, E., C. Bracis, M. Delgado, T. D. Meckley, I. Kojola and C. M. Wagner (2015). "A comparison of methods and practical guide to the behavioral analysis of animal movements." Journal of Animal Ecology Special Issue. Hoye, T. R., V. Dvornikovs, J. M. Fine, K. R. Anderson, C. S. Jeffrey, D. C. Muddiman, F. Shao, P. W. Sorensen and J. Wang (2007). "Details of the structure determination of the sulfated steroids PSDS and PADS: new components of the sea lamprey (petromyzon marinus) migratory pheromone." Journal of Organic Chemistry 72(20): 7544-7550. Johnson, N., A. Muhammad, H. Thompson, J. Choi and W. Li (2012). "Sea lamprey orient toward a source of a synthesized pheromone using odor-conditioned rheotaxis." Behavioral Ecology and Sociobiology 66(12): 1557-1567. Johnson, N. S., M. A. Luehring, M. J. Siefkes and W. M. Li (2006). "Mating pheromone reception and induced behavior in ovulating female sea lampreys." North American Journal of Fisheries Management 26(1): 88-96. Johnson, N. S., M. J. Siefkes, C. M. Wagner, H. Dawson, H. Y. Wang, T. Steeves, M. Twohey and W. M. Li (2013). "A synthesized mating pheromone component increases adult sea lamprey (Petromyzon marinus) trap capture in management scenarios." Canadian Journal of Fisheries and Aquatic Sciences 70(7): 1101-1108. Johnson, N. S., S. S. Yun, H. T. Thompson, C. O. Brant and W. M. Li (2009). "A synthesized pheromone induces upstream movement in female sea lamprey and summons them into traps." Proceedings of the National Academy of Sciences of the United States of America 106(4): 10211026. Jorgensen, J. C. and J. F. Kitchell (2005). "Growth potential and host mortality of the parasitic phase of the sea lamprey (Petromyzon marinus) in Lake Superior." Canadian Journal of Fisheries and Aquatic Sciences 62(10): 2343-2353. Kells, A. and D. Goulson (2001). "Evidence for Handedness in Bumblebees." Journal of Insect Behavior 14(1): 47-55. Li, W., M. Twohey, M. Jones and M. Wagner (2007). "Research to Guide Use of Pheromones to Control Sea Lamprey." Journal of Great Lakes Research 33, Supplement 2(0): 70-86. 294 Lohmann, K. J., C. M. F. Lohmann and C. S. Endres (2008). "The sensory ecology of ocean navigation." Journal of Experimental Biology 211(11): 1719-1728. Luehring, M. A., C. M. Wagner and W. M. Li (2011). "The efficacy of two synthesized sea lamprey sex pheromone components as a trap lure when placed in direct competition with natural male odors." Biological Invasions 13(7): 1589-1597. Meckley, T., C. Holbrook, C. Wagner and T. Binder (2014a). "An approach for filtering hyperbolically positioned underwater acoustic telemetry data with position precision estimates." Animal Biotelemetry 2(1): 7. Meckley, T., C. Wagner and E. Gurarie (2014b). "Coastal movements of migrating sea lamprey (Petromyzon marinus) in response to a partial pheromone added to river water: implications for management of invasive populations." Can J Fish Aquat Sci 71(4): 533 - 544. Meckley, T. D., C. M. Wagner and M. A. Luehring (2012). "Field evaluation of larval odor and mixtures of synthetic pheromone components for attracting migrating sea lampreys in rivers." Journal of Chemical Ecology 38(8): 1062-1069. Mellor, G. L. and T. Yamada (1982). "Development of a turbulence closure model for geophysical fluid problems." Reviews of Geophysics 20(4): 851-875. Moore, H. H. and I. P. Schleen (1980). "Changes in spawning runs of Sea Lamprey (Petromyzon-Marinus) in selected streams of Lake Superior after chemical control." Canadian Journal of Fisheries and Aquatic Sciences 37(11): 1851-1860. Morman, R. H., D. W. Cuddy and P. C. Rugen (1980). "Factors Influencing the Distribution of Sea Lamprey (Petromyzon marinus) in the Great Lakes." Canadian Journal of Fisheries and Aquatic Sciences 37(11): 1811-1826. Morshedian, A. and Gordon L. Fain (2015). "Single-Photon Sensitivity of Lamprey Rods with Cone-like Outer Segments." Current Biology 25(4): 484-487. Mullett, K. M., J. W. Heinrich, J. V. Adams, R. J. Young, M. P. Henson, R. B. McDonald and M. F. Fodale (2003). "Estimating Lake-wide Abundance of Spawning-phase Sea Lampreys (Petromyzon marinus) in the Great Lakes: Extrapolating from Sampled Streams Using Regression Models." Journal of Great Lakes Research 29, Supplement 1(0): 240-252. 295 Nguyen, T. D., P. Thupaki, E. J. Anderson and M. S. Phanikumar (2014). "Summer circulation and exchange in the Saginaw Bay-Lake Huron system." Journal of Geophysical Research: Oceans 119(4): 2713-2734. Pe'er, G., D. Saltz, H.-H. Thulke and U. Motro (2004). "Response to topography in a hilltopping butterfly and implications for modelling nonrandom dispersal." Animal Behaviour 68(4): 825839. R Development Core Team (2015). R: A language and environment for statistical computing. Vienna, Austria, R Foundation for Statistical Computing. Rao, Y. R. and D. J. Schwab (2007). "Transport and mixing between the coastal and offshore waters in the Great Lakes: a review." Journal of Great Lakes Research 33(1): 202-218. Siefkes, M. J., S. R. Winterstein and W. Li (2005). "Evidence that 3-keto petromyzonol sulphate specifically attracts ovulating female sea lamprey, Petromyzon marinus." Animal Behaviour 70: 1037-1045. Silva, S., M. J. Araújo, M. Bao, G. Mucientes and F. Cobo (2014). "The haematophagous feeding stage of anadromous populations of sea lamprey Petromyzon marinus: low host selectivity and wide range of habitats." Hydrobiologia 734(1): 187-199. Smagorinsky, J. (1963). "General circulation experiments with the primiate equations." Monthly Weather Review 91(3): 99-164. Sorensen, P. W., L. A. Vrieze and J. M. Fine (2003). "A multi-component migratory pheromone in the sea lamprey." Fish Physiology and Biochemistry 28(1-4): 253-257. Teeter, J. H. (1980). "Pheromone communication in sea lampreys (Petromyzon marinus): implications for population management." Canadian Journal of Fisheries and Aquatic Sciences 37: 2123-2132. Vickers, N. J. (2000). "Mechanisms of animal navigation in odor plumes." Biological Bulletin 198(2): 203-212. Vrieze, L., R. Bergstedt and P. Sorensen (2011). "Olfactory-mediated stream-finding behavior of migratory adult sea lamprey (Petromyzon marinues)." Can J Fish Aquat Sci 68: 523 - 533. 296 Vrieze, L. A., R. Bjerselius and P. W. Sorensen (2010). "Importance of the olfactory sense to migratory sea lampreys Petromyzon marinus seeking riverine spawning habitat." Journal of Fish Biology 76(4): 949-964. Vrieze, L. A. and P. W. Sorensen (2001). "Laboratory assessment of the role of a larval pheromone and natural stream odor in spawning stream localization by migratory sea lamprey (Petromyzon marinus)." Canadian Journal of Fisheries and Aquatic Sciences 58(12): 2374-2385. Wagner, C. M., M. B. Twohey and J. M. Fine (2009). "Conspecific cueing in the sea lamprey: do reproductive migrations consistently follow the most intense larval odour?" Animal Behaviour 78(3): 593-599. Waldman, J., C. Grunwald and I. Wirgin (2008). Sea lamprey Petromyzon marinus: an exception to the rule of homing in anadromous fishes. Zimmer-Faust, R. K., C. M. Finelli, N. D. Pentcheff and D. S. Wethey (1995). "Odor plumes and animal navigation in turbulent water flow: a field study." The Biological Bulletin 188(2): 111116. 297 CHAPTER 5 ISOLATING OPPORTUNITIES FOR MANIPULATING SPAWNING HABITAT SELECTION OF INVASIVE SEA LAMPREY (PETROMYZON MARINUS) IN THE LAURENTIAN GREAT LAKES ABSTRACT Recent advances in the understanding of the sea lamprey lake migration have shed light on how sea lamprey encounter, enter, and ascend rivers in the Great Lakes. Herein, a simple framework is provided for discussion of the factors that influence the transition between each step of the migration and how these factors could be manipulated for improved management of the invasive sea lamprey. A profitable alternative management effect either reduces the amount of pesticide applied, increases the effect of the pesticide (reduces cost per kill), or allows for the removal of barriers increasing stream continuity, while maintaining or decreasing the sea lamprey population. Six management scenarios are discussed including what behaviors would be manipulated in each scenario, what aspects of the migration need to be better understood to determine the probability of success, and what scientific advancements are necessary for each manipulation to be performed in nature. INTRODUCTION The invasive sea lamprey has been historically controlled by preventing access to habitat with dams and killing larvae with pesticides in the areas of watersheds sea lamprey could reach, driving the population towards, but not to, management targets designed to protect the ecosystem (Christie and Goddard 2003). Alternatives to traditional management options or maximization of current control methods could be valuable if they achieve sea lamprey management objectives or maintain the current efficacy with fewer negative effects. The current methods include 298 lampricides which kill non-target species (Boogaard et al. 2003) and are expensive (Christie and Goddard 2003), and physical barriers that prevent access to other migratory species. Particularly promising is the potential to use natural odorants that are part of the habitat and mate selection decisions of sea lamprey to manipulate their behavior for control (Li et al. 2007). Three elements make behavioral manipulation of a pest for protection of a resource possible: a known behavior, a known stimulus that initiates the behavior, and a scenario in which the stimulus can be presented to induce the behavior subsequently resulting in protection of a resource (Foster and and Harris 1997). Behavioral manipulation of the potadromous Great Lakes sea lamprey is promising, given their known reliance on odorants to select habitat (Teeter 1980). However, a lack of understanding of much of their life cycle and the challenge of identifying and replicating the odorants has made this alternative control method difficult to achieve. Sea lamprey are known to alter behavior in response to the odor of larval lamprey, adult sea lamprey, and dead adult or larval sea lamprey as they complete a non-homing return migration from offshore feeding grounds to riverine spawning grounds (Bjerselius et al. 2000, Siefkes 2000, Fine et al. 2004, Waldman et al. 2008, Wagner et al. 2011, Bals 2012). The odorants represent a potential tool for controlling habitat selection, trapping sea lamprey, or disrupting the migration system by masking odorants in the system through pumping synthetic replicates of the odors at high concentrations (Teeter 1980), classic concepts of insect pheromone-based control strategies (Steiner 1952, Carde and Minks 1995). Odorants can act as a releaser (e.g. a male sex pheromone component, 3kPZS, induces upstream search for mates), and (or) as a primer, (e.g., 3kPZS modulates release of gonadotropic releasing hormone, resulting in maturation of immature sea lamprey), helping to synchronize the migration (Brennan and Zufall 2006, Chung-Davidson et al. 2013). We are focused on releaser effects for manipulating 299 behavior. Though progress has been made in understanding the migration system and identification of some odorants, an understanding of how to best use odor in a control scenario has not materialized. Through a simple migration model that is analogous in some respects to trap catch steps previously developed for sea lamprey (Bravener and McLaughlin 2013), we identify the steps of the migration, although in this case the sea lamprey reaches spawning habitat instead of entering a trap and we identify the steps of the migration, the factors that influence transitions, and how the process can be manipulated. RIVER ENTRY FRAMEWORK The framework covers river plume encounter, river entry, and river ascension to the spawning grounds. It allows us to outline what factors influence each step and to discuss the potential to manipulate these steps to better manage sea lamprey in the Great Lakes (Figure 5.1). River Plume Encounter The number of sea lamprey to encounter a river plume will depend on the number of individuals that pass the river mouth and how far they are from shore compared to the distance the river plume extends from shore (Figure 5.2). The probability of encounter for each river will vary with the size of the river plume, principally how far river water extends from the coast and how far from shore sea lamprey traverse parallel to the coast in that region (Chapter 4). River plume size varies with river discharge, coastal bathymetry, coastline configuration at the river mouth (e.g., embayment trapping water), and wind driven lake current (Churchill et al. 2003, Rao and Schwab 2007). Wind driven currents and the coastal boundary are the dominant features regulating the extent of the river plume in the Great Lakes (Churchill et al. 2003). The number of sea lamprey available to encounter the plume will depend on the number of sea lamprey to return 300 to the particular region following the parasitic stage and whether they reach a given river or stop at a previous river. River water alone is attractive to sea lamprey in two-choice tests and appears to be the primary feature that initiates a transition to local search behavior around the river mouth in the lake (Vrieze and Sorensen 2001, Vrieze et al. 2011, Meckley et al. 2014, Chapter 4). However nose plugged sea lamprey were still able to locate a nearby river, albeit much less efficiently (12.5 %) (Vrieze et al. 2010), indicating that other features such as conspecific attraction or physical features such as substrate, water color (tannins), temperature gradients, or outflow may be sufficient for a migrant to recognize their proximity to a river mouth and enter the river. River Entrance, Upstream Movement, and Tributary Selection The timing of river entry and upstream movement is related to river temperature, changes in river temperature, stream flow, and sexual maturation (Almeida et al. 2002, Binder et al. 2010). River and tributary selection appears most reliant on the presence of larval sea lamprey odor, a form of public information that provides evidence of spawning and rearing habitat quality (Bjerselius et al. 2000, Sorensen et al. 2005, Waldman et al. 2008). However, additional odorants influence sea lamprey behavior and may influence the migration. For example, it is unclear how public information such as larval odor of other lamprey species or decaying lamprey larvae modulates river entry. Unlike specialized sex pheromones that are often a compilation of a few compounds at a specific ratio (Ando et al. 2004), tuned as a result of receiver bias, when an odor releaser is rewarded with more mates because of releasing a more attractive odor (Buchinger et al. 2013); the unspecialized release of larval odor is unlikely to be discernable as a species specific signal by adults and may be a larger number of components (Li et al. 1995). The composition of the 301 gut microbiome of sea lamprey is just beginning to be explored (Tetlock et al. 2012), but if the larval diet, gut microbiome, and subsequent bile salts released by larvae are similar between different types of larvae, larval odor may be a general lamprey habitat quality cue. This is supported by evidence that different species of lamprey showed attraction to different lamprey species larval odors and larvae of different species release similar components (Fine et al. 2004). Although a blend of active compounds must be present (Fine and Sorensen 2008, Meckley et al. 2012), odors are released to a constrained environment and ratios may not be important (Derby and Sorensen 2008). Little is known about the influence of the odor of dead larvae on the sea lamprey migration, beyond that the odor of dead lamprey larvae is repulsive to migratory sea lamprey in lab tests (Imre et al. 2010, Wagner et al. 2011, Bals and Wagner 2012), and has been hypothesized to cause habitat avoidance where larval sea lamprey have died during the winter. Migratory sea lamprey avoid the odor of dead adults and larvae in the lab, though it is unclear if different compounds are responsible for the avoidance (Bals and Wagner 2012). These odor cues are unlikely to be species specific. The second type of information widely relied upon by animals but poorly described in sea lamprey is reliance on interactions with mates, known as conspecific attraction (Doligez et al. 2003). The conspecific attraction strategy always coexists with other strategies because it allows individuals to utilize the decisions of other individuals and ensures the availability of mates (Doligez et al. 2003). The presence of mates could be determined visually or through olfaction as sea lamprey migrants can smell the odor of other migrants at different stages of the migration and the odor of dead mates upstream. To date the natural progression of the sea lamprey migration has been poorly observed as studies utilize subjects trapped in a river (Vrieze et al. 2010, Vrieze et al. 2011, Meckley et al. 2014 ), and activities like staging at a river mouth could 302 be a major component of the migration (Applegate 1950, Clemens et al. 2010), as it could signal the presence of mates and synchronize the start of the migration. Unlike larval odors these cues are more likely to be species specific migration cues. The odor of adult conspecifics synchronizes adult maturation in the river (ChungDavidson et al. 2013) and the presence of a male released sex pheromone may encourage female river entry (Meckley et al. 2014) and tributary selection. However, sex pheromone was not found to be attractive to immature migratory sea lamprey caught in early spring (Siefkes et al. 2005), and may never be available at a river mouth to migrants experiencing a natural migration progression. More progress has been made towards synthetically replicating the sex pheromone (Johnson et al. 2005), though an effective large scale management scenario has not been identified to date (Luehring et al. 2011, Johnson et al. 2013) and as a trap bait where an effect has been shown, the method may only be effective at low sea lamprey densities (El-Sayed et al. 2006, Johnson et al. 2013). The odor of dead migratory sea lamprey is repulsive to non-mature male and female sea lamprey in lab tests (Imre et al. 2010, Wagner et al. 2011, Bals and Wagner 2012) and may represent predator avoidance or the end of spawning, indicating sea lamprey have already reproduced and senesced. The function of dead odor is further confounded by the finding that migrants become habituated to the odor of dead adults after approximately four hours of exposure (Bals 2012), and initial field tests suggest that the odor of dead sea lamprey is insufficient to prevent river entry and instead increased river entry (Tom Luhring, Personal Communication). The odor functions to bias the side of the stream channel occupied by sea lamprey during upstream movement when the odor can be avoided (Bals 2012), but may increase 303 upstream movement past the source of the odor when the odor is mixed throughout the entire channel (Tom Luhring, Personal Communication). River Retention If a river contains a large larval population, spawning habitat, and conspecific mates, sea lamprey are unlikely to reverse course after river entrance. This is supported by a two-year study that never observed sea lamprey exiting a river after they entered and moved more than 300 m upstream, and some individuals enter the immediate river mouth without ascending the river (Applegate 1950, Meckley et al. 2014, Chapter 4), together leading to the suggestion that the decision to enter occurs at the river mouth (Chapter 4). Alternatively, some sea lamprey return to a river plume over multiple nights without ever entering the river, regularly reversing between movement inside and outside of the river plume, and moving very close to the river mouth, although this occurred less with higher larval odor (Chapter 4). Observations that sea lamprey moved to the river mouth without entering supports that the concept that the decision to enter a river usually occurs at the river mouth. MANIPULATING THE MIGRATION Management of an invasive species is difficult because a strong understanding of the species biology and behavior in the new environment is often unavailable. Because eradication of established invasive species is rare (Simberloff 2003), an explicit definition of what constitutes a profitable outcome is required. We define a profitable alternative management effect as one that reduces the amount of pesticide applied, increases the effect of the pesticide (reduces cost per kill), or allows for the removal of barriers, while maintaining or decreasing the sea lamprey population. The results must be measureable to determine whether a strategy has a profitable effect, a difficult proposition given the dynamic pesticide treatment schedule. An 304 effective odor-based control method must identify and synthesize the active compounds cheaply, a very elusive process (Li et al. 2007), that we will not discuss in detail. We discuss six management scenarios for manipulating the sea lamprey migration for improved management (Table 5.1). River Plume Encounter Encounter with a river plume must occur for a sea lamprey to enter the river. River plume encounter could be manipulated by changing the number of individuals to pass the river, physically influencing the distance that sea lamprey move from the coast while passing the river plume, through providing chemical information indicative of a river plume farther from shore, or by physically increasing the distance the river plume extends into the lake. Physically altering the distance sea lamprey travel from shore or the size of the river plume would likely require large construction investments and extensive research. However, there are navigational features that may provide suggestions for how manipulation of the river plume size could be achieved (Figure 5.3). An alternative would be the release of river odor out in front of a river mouth that indicates the presence of a nearby river, inducing local search (Table 5.1, Option 1). The encounter rate may be challenging to alter; thereby changing the number of sea lamprey to pass a river is the most plausible approach if entry rates of nearby rivers can be influenced. River plume encounter is an event not a decision. For this reason it may not have management value unless encounter can be prevented, however if it is combined with some decision manipulation following encounter, the encounter rate becomes important. Cohort-Size Maintenance If synthetic larval odor was available, the most immediate value could come in the form of maintaining the number of sea lamprey to enter streams that require pesticide treatment every 305 four years (Table 5.1, Option 2; Figure 5.2, iii). If the larval odor signal is maintained, migrant dispersion could be abated and could reduce the cohort size of sea lamprey to enter other rivers, potentially including small river plumes that may get irregular sea lamprey recruitment. It is unconfirmed whether the increased river entry rates observed each year after treatment is due only to the increase in larval odor, or if for example an increase in river recruitment of 2.4-6.8% of the total population of Lake Huron (i.e., as observed between 1 and 3 years post treatment in the Ocqueoc River; Figure 5.4), is sufficient for a lake wide management value. A potential downside to the strategy is the need for continued pesticide treatments. There are many smaller streams that have sea lamprey populations that are too small, or are too dispersed to be treated effectively and never or rarely meet the standard to qualify for treatment (Jones et al. 2003). This strategy works under the premise that the constant presence of larval odor in rivers with large river plumes that are regularly treated would reduce the spreading of migrants to other rivers by reducing the coastal search time and associated likelihood of encountering smaller rivers. This strategy is only important if sea lamprey are spread to non-expert treatment streams that do not get treated every four years. If sea lamprey are only being pushed between regularly treated streams, then the cohort-size maintenance approach is not valuable. Recruitment to habitat deficient for larval survival Synthetic larval odor could encourage entry into rivers with poor spawning or rearing habitat (Li et al. 2007; Table 5.1, Option 3). We suspect a river with limited habitat, such as a river with a physical barrier to movement low in the river before spawning habitat is reached, is unlikely to retain migrants, as sea lamprey have been observed exiting large rivers with waterfalls low in the river (Teeter 1980). However if a river has ample spawning habitat and poor larval habitat, sea lamprey may stay and spawn. It would need to be confirmed that larvae 306 do survive poorly in any river site considered for this strategy. Encouraging sea lamprey into a river lacking larvae or with few larvae could be counterproductive if enough larvae survive to warrant pesticide treatment of the river or if it results in an unacceptable number of parasites. It is unclear if there are rivers that fit this strategy as most rivers that do not contain larvae, are intermittent or have very low flow (Teeter 1980). Chemical barrier and Push-Pull The potential to create a chemical barrier at a river mouth or at tributaries has begun to be tested (Bals 2012, Bals and Wagner 2012, C. Michael Wagner personal communication; Figure 5.2, iv). Currently only physical barriers are plausible for blocking sea lamprey access to rivers. Dead adult odor does not prevent river entry, though the dead odor of larvae has not been tested in this scenario (Table 5.1, Option 6). Parts of a watershed may be blocked through application of dead migratory sea lamprey odor alone, although the application method will need to avoid extended exposure resulting in odor habituation which was observed after sea lamprey were exposed to the odor for approximately four hours (Bals 2012, Bals and Wagner 2012, Tom Luehring personal communication). The effect may be strengthened with a “pull” comprised by larval odor, as sea lampreys prefer to move in the stronger larval signal (Wagner et al. 2009, Meckley et al. 2012; Table 5.1, Option 4; Figure 5.2, v, vi). This method of providing a chemical barrier to part of the watershed has promise to limit sea lamprey to a narrow region of the watershed that could reduce the pesticide necessary for treatment. The level of success necessary for a management value would need to be evaluated and the method will fail to manipulate individuals that spawn before encountering the selected stream bifurcation. If sea lamprey are sufficiently repelled to delay the need for pesticide 307 treatment, or allow for a physical barrier to be removed or placed higher in the watershed, the method would have value. Larval Odor Removal (TFM) We only highlight the use of TFM as an alternative control measure because the way it is applied removes the larval signal from a river and has a measurable effect on entry into neighboring rivers (e.g., Teeter 1980; Table 5.1, Option 5). The more effective the removal of lamprey larvae in the river the more effectively the treatment will likely reduce entry the following year, creating a repellent from the absence of a cue. In the circumstance pesticide treatment occurs in many rivers, the reduction in entry by migrants the year after treatment is a negative management effect. If TFM treatment is required every four years before larvae can transform to the parasitic stage, no management advantage is gained from the reduction in entry and instead sea lamprey are spread to other rivers. Although where migrants are dispersed the year after treatment, is poorly understood. In many cases, only a few rivers receive a large proportion of the sea lamprey population estimated in the lake, removing the larval signal from one of these rivers can cause a substantial shift in where migrants go to spawn and the more time spent searching may increase the likelihood of encountering smaller river plumes that are rarely encountered (Meckley et al. 2014). Based on similar logic, we would recommend against the treatment of multiple large rivers in a region in the same year. Ideally alternating pesticide application between large rivers may result in shifting recruitment between the major rivers rather than encouraging encounter and entry with many smaller rivers, a concept that needs further testing. Similarly we recommend against concepts like treatment of all rivers in back to back years, as this is likely to result in spreading migrants broadly. Encouraging dispersal of a pest is counterproductive to 308 chemical based treatment. Finally shifting treatments to after the spawning season would be the most valuable as the larval signal that acts to congregate migrants would congregate sea lamprey in rivers that were about to be treated. It is likely infeasible to treat all rivers after spawning occurs although the more the treatments can be shifted to after the spawning season the more effectively the treatment program would be using the natural sea lamprey habitat selection decisions to the control program’s advantage. CONCLUSION The importance of pesticide application and physical barriers to sea lamprey movement will not be quickly replaced. As knowledge of the sea lamprey migration increases more options become available for how the sea lamprey migration can be manipulated. There are variants of these six potential alternative management strategies, although they cover the primary aspects of the migration that can be manipulated (Figure 5.1; Figure 5.2). 309 APPENDIX 310 Table 5.1: Six general strategies for manipulating sea lamprey migration behavior are presented including the stimulus affected, how the stimulus is affected (action), the goal of the manipulation and the type of manipulation. Only option one does not stand alone as a management strategy as it manipulates behavior but not a decision to select habitat. Stimulus Altered Manipulation Action Intended Result Type of Manipulation 1 River water Release river water out in front of a river mouth River Plume Assessment Pull/transition 2 Larval Odor Add larval odor to “expert judgement” stream River Entry Pull 3 Larval Odor Add larval odor to stream with poor larval survival River Entry Pull 4 Larval Odor And Dead Odor Add larval odor and dead lamprey odor Tributary selection Push-Pull 5 Larval Odor Remove larval odor signal via lampricide application (TFM) River rejection Push 6 Dead Larval Odor Add dead lamprey odor to entire river River avoidance Push 311 Figure 5.1: The steps of the spawning migration are analogous to the steps involved in trapping sea lamprey (Bravener and McLaughlin 2013), although in this case the river takes the place of the trap and the model depicts the transitions that occur when a sea lamprey enters a river in the great lakes. The arrows refer to the probabilities of progressing from the lake to spawning habitat or reversing through any prior step. The P refers to the probability of the event. 312 Figure 5.2: The distance sea lamprey travel from shore (a, dotted arrow), the distance the river plume extends from the coast (b), the presence of other rivers on the coast and the number of rivers encountered prior to reaching a river (c), are all important to a sea lamprey encountering a particular river plume (Region of Coast). Following encounter with a river plume (d) sea lamprey become 313 Figure 5.2: (cont’d) available for manipulation of their habitat selection decisions including river entry (e), or tributary entry (f). Five general manipulation scenarios are shown including the manipulation of river plume encounter (event) (ii) or habitat selection (decision) (iiivi). The operator (+, - , =), indicates how the manipulation of adding an attractant (“attract”), a repellant, or removing an attractant (“repel”), influences encounter rates at nearby rivers and entry of those rivers. The manipulation of decisions through the addition of odorants (iii-vi), uses general terms based on the effect desired and not specific odorants that will cause the desired effect. Encounter is shown to change due to the order of sequential river plumes. Encounter at a river can change due to a manipulation at a prior river but not the current river manipulated (decrease, iii; increase, iv). It is unclear how past experience might influence entry of future rivers (iii, “?”). 314 Figure 5.3: Many river mouths have preexisting structures that may extend the distance the river plume would otherwise reach from shore (i.e., as seen west of river mouths in Ludington, MI (left) and Sagatuck, MI (right)). Structures of this type designed for boat traffic could present unique solutions for influencing the sea lamprey migration and may be the only scenario that would create a river plume extending a greater distance from shore than would otherwise occur for that river. In each image the river plume can be faintly seen due to darker turbidity. 315 Figure 5.4: The percentage of the entire population of adult sea lamprey in Lake Huron that return to the Ocqueoc River to spawn from 1995 to 2015 based on mark recapture (black circles) and the change in entry between years (blue circles). A change of entry rate of 2 represents a doubling in the rate of entry. Vertical lines represent the times of TFM treatment to the Ocquoec River. The change in entry rate from one year after TFM treatment to the second year shows that the increase in entry rate between 2010 and 2011 (D) fits within the historical increase in rate (A: 1.8, B: 1.3, C: 3.1, D: 2.1). 316 REFERENCES 317 REFERENCES Almeida, P. R., B. R. Quintella and N. M. Dias (2002). "Movement of radio-tagged anadromous sea lamprey during the spawning migration in the River Mondego (Portugal)." Hydrobiologia 483(1-3): 1-8. Ando, T., S.-i. Inomata and M. Yamamoto (2004). Lepidopteran Sex Pheromones. The Chemistry of Pheromones and Other Semiochemicals I. S. Schulz, Springer Berlin Heidelberg. 239: 51-96. Applegate, V. C. (1950). The natural history of the sea lamprey in Michigan. Washington, D. C, U. S. Department of Interior Fish & Wildlife Service, Special Scientific Report Fisheries. Bals, J. D. (2012). The sea lamprey alarm response: field and laboratory investigations. M.S., Michigan State Unviersity. Bals, J. D. and C. M. Wagner (2012). "Behavioral responses of sea lamprey (Petromyzon marinus) to a putative alarm cue derived from conspecific and heterospecific sources." Behaviour 149(9): 901-923. Binder, T. R., R. L. McLaughlin and D. G. McDonald (2010). "Relative Importance of Water Temperature, Water Level, and Lunar Cycle to Migratory Activity in Spawning-Phase Sea Lampreys in Lake Ontario." Transactions of the American Fisheries Society 139(3): 700-712. Bjerselius, R., W. Li, J. H. Teeter, J. G. Seelye, P. B. Johnson, P. J. Maniak, G. C. Grant, C. N. Polkinghorne and P. W. Sorenson (2000). "Direct behavioral evidence that unique bile acids released by larval sea lamprey (Petromyzon marinus) function as a migratory pheromone." Canadian Journal of Fisheries and Aquatic Sciences 57: 557-569. Boogaard, M. A., T. D. Bills and D. A. Johnson (2003). "Acute toxicity of TFM and a TFM/niclosamide mixture to selected species of fish, including lake sturgeon (Acipenser fulvescens) and mudpuppies (Necturus maculosus), in laboratory and field exposures." Journal of Great Lakes Research 29: 529-541. Bravener, G. A. and R. L. McLaughlin (2013). "A behavioural framework for trapping success and its application to invasive sea lamprey." Canadian Journal of Fisheries and Aquatic Sciences 70(10): 1438-1446. 318 Brennan, P. A. and F. Zufall (2006). "Pheromonal communication in vertebrates." Nature 444(7117): 308-315. Buchinger, T. J., H. Wang, W. Li and N. S. Johnson (2013). Evidence for a receiver bias underlying female preference for a male mating pheromone in sea lamprey. Carde, R. T. and A. K. Minks (1995). "Control of moth pests by mating disruption: successes and constraints." Annual review of entomology 40(1): 559-585. Christie, G. C. and C. I. Goddard (2003). "Sea Lamprey International Symposium (SLIS II): Advances in the integrated management of sea lamprey in the Great Lakes." Journal of Great Lakes Research 29: 1-14. Chung-Davidson, Y. W., H. Y. Wang, M. J. Siefkes, M. B. Bryan, H. Wu, N. S. Johnson and W. M. Li (2013). "Pheromonal bile acid 3-ketopetromyzonol sulfate primes the neuroendocrine system in sea lamprey." Bmc Neuroscience 14: 13. Churchill, J. H., E. A. Ralph, A. M. Cates, J. W. Budd and N. R. Urban (2003). "Observations of a negatively buoyant river plume in a large lake." Limnology and Oceanography 48(2): 884-894. Clemens, B. J., T. R. Binder, M. F. Docker, M. L. Moser and S. A. Sower (2010). "Similarities, Differences, and Unknowns in Biology and Management of Three Parasitic Lampreys of North America." Fisheries 35(12): 580-594. Derby, C. and P. Sorensen (2008). "Neural Processing, Perception, and Behavioral Responses to Natural Chemical Stimuli by Fish and Crustaceans." Journal of Chemical Ecology 34(7): 898914. Doligez, B., C. Cadet, E. Danchin and T. Boulinier (2003). "When to use public information for breeding habitat selection? The role of environmental predictability and density dependence." Animal Behaviour 66(5): 973-988. El-Sayed, A. M., D. M. Suckling, C. H. Wearing and J. A. Byers (2006). Potential of Mass Trapping for Long-Term Pest Management and Eradication of Invasive Species. Fine, J., L. Vrieze and P. Sorensen (2004). "Evidence That Petromyzontid Lampreys Employ a Common Migratory Pheromone That Is Partially Comprised of Bile Acids." Journal of Chemical Ecology 30(11): 2091-2110. 319 Fine, J. M. and P. W. Sorensen (2008). "Isolation and biological activity of the multi-component sea lamprey migratory pheromone." Journal of chemical ecology 34(10): 1259-1267. Foster and, S. P. and M. O. Harris (1997). "Behavioral manipulation methods for insect pestmanagement." Annual Review of Entomology 42(1): 123-146. Imre, I., G. E. Brown, R. A. Bergstedt and R. McDonald (2010). "Use of Chemosensory Cues as Repellents for Sea Lamprey: Potential Directions for Population Management." Journal of Great Lakes Research 36(4): 790-793. Johnson, N. S., M. J. Siefkes and W. M. Li (2005). "Capture of ovulating female sea lampreys in traps baited with spermiating male sea lampreys." North American Journal of Fisheries Management 25(1): 67-72. Johnson, N. S., M. J. Siefkes, C. M. Wagner, H. Dawson, H. Y. Wang, T. Steeves, M. Twohey and W. M. Li (2013). "A synthesized mating pheromone component increases adult sea lamprey (Petromyzon marinus) trap capture in management scenarios." Canadian Journal of Fisheries and Aquatic Sciences 70(7): 1101-1108. Jones, M. L., R. A. Bergstedt, M. B. Tvohey, M. F. Fodale, D. W. Cuddy and J. W. Slades (2003). "Compensatory mechanisms in great lakes sea lamprey populations: Implications for alternative control strategies." Journal of Great Lakes Research 29: 113-129. Li, W., M. Twohey, M. Jones and M. Wagner (2007). "Research to Guide Use of Pheromones to Control Sea Lamprey." Journal of Great Lakes Research 33, Supplement 2(0): 70-86. Li, W. M., P. W. Sorensen and D. D. Gallaher (1995). "The olfactory system of migratory adult sea lamprey (Petromyzon marinus) is specifically and acutely sensitive to unique bile-acids released by conspecific larvae." Journal of General Physiology 105(5): 569-587. Luehring, M. A., C. M. Wagner and W. M. Li (2011). "The efficacy of two synthesized sea lamprey sex pheromone components as a trap lure when placed in direct competition with natural male odors." Biological Invasions 13(7): 1589-1597. Meckley, T., C. Wagner and E. Gurarie (2014). "Coastal movements of migrating sea lamprey (Petromyzon marinus) in response to a partial pheromone added to river water: implications for management of invasive populations." Can J Fish Aquat Sci 71(4): 533 - 544. 320 Meckley, T. D., C. M. Wagner and M. A. Luehring (2012). "Field evaluation of larval odor and mixtures of synthetic pheromone components for attracting migrating sea lampreys in rivers." Journal of Chemical Ecology 38(8): 1062-1069. Rao, Y. R. and D. J. Schwab (2007). "Transport and mixing between the coastal and offshore waters in the Great Lakes: a review." Journal of Great Lakes Research 33(1): 202-218. Siefkes, M. J. (2000). Spermiating male sea lampreys release a sex pheromone that attracts postovulatory female sea lampreys. Masters, Michigan State University. Siefkes, M. J., S. R. Winterstein and W. Li (2005). "Evidence that 3-keto petromyzonol sulphate specifically attracts ovulating female sea lamprey, Petromyzon marinus." Animal Behaviour 70: 1037-1045. Simberloff, D. (2003). "How Much Information on Population Biology Is Needed to Manage Introduced Species?" Conservation Biology 17(1): 83-92. Sorensen, P. W., J. M. Fine, V. Dvornikovs, C. S. Jeffrey, F. Shao, J. Wang, L. A. Vrieze, K. R. Anderson and T. R. Hoye (2005). "Mixture of new sulfated steroids functions as a migratory pheromone in the sea lamprey." Nature Chemical Biology 1(6): 324-328. Steiner, L. F. (1952). "Methyl eugenol as an attractant for oriental fruit fly." Journal of Economic Entomology 45(2): 241-248. Teeter, J. H. (1980). "Pheromone communication in sea lampreys (Petromyzon marinus): implications for population management." Canadian Journal of Fisheries and Aquatic Sciences 37: 2123-2132. Tetlock, A., C. K. Yost, J. Stavrinides and R. G. Manzon (2012). "Changes in the Gut Microbiome of the Sea Lamprey during Metamorphosis." Applied and Environmental Microbiology 78(21): 7638-7644. Vrieze, L., R. Bergstedt and P. Sorensen (2011). "Olfactory-mediated stream-finding behavior of migratory adult sea lamprey (Petromyzon marinues)." Can J Fish Aquat Sci 68: 523 - 533. Vrieze, L. A., R. Bjerselius and P. W. Sorensen (2010). "Importance of the olfactory sense to migratory sea lampreys Petromyzon marinus seeking riverine spawning habitat." Journal of Fish Biology 76(4): 949-964. 321 Vrieze, L. A. and P. W. Sorensen (2001). "Laboratory assessment of the role of a larval pheromone and natural stream odor in spawning stream localization by migratory sea lamprey (Petromyzon marinus)." Canadian Journal of Fisheries and Aquatic Sciences 58(12): 2374-2385. Wagner, C. M., E. M. Stroud and T. D. Meckley (2011). "A deathly odor suggests a new sustainable tool for controlling a costly invasive species." Canadian Journal of Fisheries and Aquatic Sciences 68(7): 1157-1160. Wagner, C. M., M. B. Twohey and J. M. Fine (2009). "Conspecific cueing in the sea lamprey: do reproductive migrations consistently follow the most intense larval odour?" Animal Behaviour 78(3): 593-599. Waldman, J., C. Grunwald and I. Wirgin (2008). Sea lamprey Petromyzon marinus: an exception to the rule of homing in anadromous fishes. 322