High-Dynamic Range Collision Detection using Piezoelectric Polymer Films for Planar and Non-planar Applications by James Michael Wooten A thesis submitted to the Graduate Faculty of Auburn University in partial ful llment of the requirements for the Degree of Master of Science Auburn, Alabama August 3, 2013 Keywords: piezoelectric, polymer, collision detection, PVDF, robotics Copyright 2013 by James Michael Wooten Approved by David M. Bevly, Co-Chair, Professor of Mechanical Engineering John Hung, Co-Chair, Professor of Electrical and Computer Engineering Dan Marghitu, Professor of Mechanical Engineering Abstract This thesis develops a large area collision detection system utilizing the piezoelectric ef- fect of polyvinylidene uoride lm. Complex high speed autonomous articulations associated with modern large-scale high degree-of-freedom (DOF) robotic arms have a high possibility of collision when integrated into human cooperative environments for human-aid, task au- tomation, and biomedical interfacing. The proposed system provides high dynamic range for sensation and robust adaptability to achieve collision detection on complex surfaces in order to augment robotic systems with collision perception. The design allows for increased cohabi- tation of human and high DOF robotic arms in cooperative environments requiring advanced and robust collision detection systems capable of retro tting onto deployed and operating robotic arms in the commercial world. Sensor testing is accomplished using multiple collision stimuli to mimic real world performance as well as impact force modeling utilizing high speed cameras. The experimentation results show a wide dynamic sensing range for collision force, from 5 N to 300 N and consistent sensor response for planar and non-planar applications. The thesis will show and support the sensor capability of wide range of collision detection while maintaining adaptability of sensor design to multiple scenarios. The approach di ers from current work which primarily focuses on small-range low levels of tactician perception, small area sensor requiring complex construction, and associated electronics and processing complexity for common approaches. The pseudo-membrane design eliminates the construc- tion complexity and limited application scope while achieving high and low levels of collision detection utilizing simple electronics and processing method. The captured experimentation results highlight the consistency of response for multiple applications, standard deviation of results less than 1 GPA, and the large range of collision detection capability from 5 N to 300 N. ii Acknowledgments The author would like to thank Siemens AG for sponsoring the research. Also, the assistance of Dr. Marghitu and Hamid Ghaednia in the e orts of impact and force modeling of the sensor. Likewise, the author would like to extend the sincerest gratitude and thanks to the GPS and Vehicle Dynamics Laboratory at Auburn University for the extensive help wanted and unwanted through out this process. As well, the author expresses appreciation for the guidance of the advisors, Dr. David Bevly and Dr. John Hung, academically, socially and spiritually over these last few years. Finally, the author would like to thank his wife and family for the logistical support in this academic marathon. iii Table of Contents Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix List of Abbreviations and Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . x 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 Contribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 2.1 Piezoelectricity and Ferroelectrics . . . . . . . . . . . . . . . . . . . . . . . . 4 2.2 Polyvinylidene Fluoride . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.3 Collision Sensing Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 2.3.1 Estimation and Control . . . . . . . . . . . . . . . . . . . . . . . . . 6 2.3.2 Collision Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.4 Hazard Classi cation and Concerns . . . . . . . . . . . . . . . . . . . . . . . 8 2.5 Motivation and Application . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.6 Coordinate System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3 Design and Modeling of the Collision Sensor . . . . . . . . . . . . . . . . . . . . 13 3.1 Piezoelectric Polymer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.2 Piezoelectric E ect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.3 Membrane Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 3.4 Pressure Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.5 Sensor Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 iv 3.6 Response Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 3.7 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 3.7.1 Ampli er Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 3.7.2 Sampling and Detection Electronics . . . . . . . . . . . . . . . . . . . 25 3.8 Design Attributes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 4 Experimentation Methods and Prototype . . . . . . . . . . . . . . . . . . . . . . 27 4.1 Prototype Sensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 4.2 Prototype Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 4.3 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 4.3.1 Collision Simulation and Approximate Force . . . . . . . . . . . . . . 29 4.3.2 Force Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 5.1 Average Force Experimentation Method . . . . . . . . . . . . . . . . . . . . 33 5.1.1 Planar Sensor Prototype . . . . . . . . . . . . . . . . . . . . . . . . . 33 5.1.2 Non-planar Sensor Prototype . . . . . . . . . . . . . . . . . . . . . . 36 5.1.3 Low Force Dynamic Response . . . . . . . . . . . . . . . . . . . . . . 38 5.1.4 Total Results Comparison and Statistical Analysis . . . . . . . . . . . 39 5.2 Force Characterized Collision Response Method . . . . . . . . . . . . . . . . 40 5.2.1 Gravity Con rmation . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 5.2.2 Force Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 5.2.3 Sensor Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 6.1 Results Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 Appendix A Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 A.1 Cauchy Stress Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 v A.2 Engineering Strain Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 A.3 Geometric Representation of Strain . . . . . . . . . . . . . . . . . . . . . . . 58 A.4 Young?s Modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 A.5 Shear Modulus of Elasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 vi List of Figures 2.1 Typical application for sensor design in non-structured environment. The shaded green area on the robotic arm represents desired collision detection areas. . . . . 10 2.2 Diagram showing Coordinate System used in Thesis . . . . . . . . . . . . . . . . 11 3.1 Membrane Sensor Construction Diagram . . . . . . . . . . . . . . . . . . . . . . 16 3.2 Pressure Sensor Construction Diagram . . . . . . . . . . . . . . . . . . . . . . . 18 3.3 Pseudo-Membrane Sensor Construction Diagram . . . . . . . . . . . . . . . . . 19 3.4 Charge Ampli er Circuit Design . . . . . . . . . . . . . . . . . . . . . . . . . . 24 4.1 Sensor Prototype used in testing showing planar (top of cover) and non-planar (rounded left end) sensor applications on example robotic arm shielding. Wires in picture connect sensor electrodes to instrumentation. . . . . . . . . . . . . . . 28 4.2 Experimental setup for force measurement of impacts. . . . . . . . . . . . . . . 32 5.1 Mean captured results for wide dynamic range of collision stimuli for planar application of sensor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 5.2 Relation of measured collision to approximate collision force. . . . . . . . . . . . 35 5.3 Mean captured results for wide dynamic range of collision stimuli. . . . . . . . . 36 5.4 The relation of measured collision to force of object collision for non-planar sensor. 37 vii 5.5 Mean Digital Response results of Element 1 and Element 2 for 5N collisions. . . 38 5.6 Measured Displacement for 5 cm. . . . . . . . . . . . . . . . . . . . . . . . . . . 40 5.7 Gravity measurements for Validity . . . . . . . . . . . . . . . . . . . . . . . . . 42 5.8 Peak Impact Force vs Height of Object . . . . . . . . . . . . . . . . . . . . . . . 43 5.9 Sensor response for 5 cm drop. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 5.10 Sensor response for 10 cm drop with responses labeled by data run notation of [N9 Date Run#]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 5.11 Sensor response for 15 cm drop with responses labeled by data run notation of [N9 Date Run#]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 5.12 Sensor Response for 20 cm drop with responses labeled by data run notation of [N9 Date Run#]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 5.13 Sensor response for 35 cm drop with responses labeled by data run notation of [N9 Date Run#]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 5.14 Sensor response for 50 cm drop. . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 5.15 Response vs Force for Total Data Run . . . . . . . . . . . . . . . . . . . . . . . 50 viii List of Tables 3.1 Material Properties of PVDF Film . . . . . . . . . . . . . . . . . . . . . . . . . 15 4.1 Collision Stimuli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 4.2 Collision Stimuli for Displacement Measurement Tests . . . . . . . . . . . . . . 31 5.1 Total results of Planar and Non-planar tests. . . . . . . . . . . . . . . . . . . . . 39 5.2 Total results of force characterized collision responses. . . . . . . . . . . . . . . 51 ix List of Abbreviations and Nomenclature Strain Shear Strain Stress Shear Stress E Young?s Modulus G Shear Modulus of Elasticity Q Charge ADC Analog to Digital Converter ak daton DOF Degree-of-freedom PVDF Polyvinylidene Flouride x Chapter 1 Introduction Complex high speed autonomous articulations associated with modern large-scale high degree-of-freedom (DOF) robotic arms have a high possibility of collision when integrated into human cooperative environments for human-aid, task automation, and biomedical in- terfacing. This thesis details a large area collision detection system utilizing the piezoelectric e ect of polyvinylidene uoride lm. The design, testing, and results in the following chap- ters will involve a novel sensor design approach and simple electronics based collision sensing solution for complex planar and non-planar surfaces in non-standard working environments. The sensor design provides high dynamic range for sensation and robust adaptability to achieve collision detection on complex surfaces in order to augment robotic systems with collision perception. The design and electronics presented allow for increased cohabitation of human and high DOF robotic arms in cooperative environments requiring advanced and robust collision detection systems capable of retro tting onto deployed and operating robotic arms in the commercial world. The majority of sensing applications rely on a lm membrane [1] or pressure [2]. A membrane approach leads to practical concerns for large-area and high impact applications because of physical limitations and a high level of design complexity for large sensor applica- tions and networks. Pressure based designs require more complex electronics and construc- tion to achieve exibility because of the low signal response and need for a sti substrate to achieve pressure dynamics. In this thesis, a novel sensor design is suggested utilizing a exible substrate to allow the lm to operate in a pseudo-membrane con guration that can achieve high dynamic range, simpli ed electronics, and robust applicability. Sensor robust- ness means that the sensor can operate in a wide variety of environments including but not 1 limited to high and low impacts, non-uniform and complex surfaces, mobile and stationary systems, and human and non-human inhabited. The proposed sensor construction allows for complex shape and non-planar surface applications. The design is intended for safety and control applications related to human-robotics interaction in cooperative environments, arm autonomy in high DOF arms in changing scenarios, and technology redundancy to minimize risk related to collision. Current safety standards limit the amount of force that a robot can impart to a human being as 150N [3]. The proposed sensor provides a dynamic sensing range of 5N to greater than 200N in an e ort to detect a state of collision before signi cant force has been imparted to the object. 1.1 Thesis Outline This thesis will discuss the design and testing of a pseudo-membrane collision detection sensor utilizing the piezoelectric e ect of PVDF lm. Some initial information will be pre- sented to give the reader a more thorough understanding of the subject matter. The following chapters of the thesis will progress through the design and modeling of the collision sensor, detail the experimental prototypes built and methods for testing, discuss and present the re- sults from experimentation, and nally conclude the work stating the contributions as well as future work possibilities. The sensor design detailing the polymer selection and construction method will be presented including reference to current sensing designs and drawbacks to available methods. Additionally, the sensor response will be modeled and physical equations presented to convey to the reader an expectation of the sensor functionality. The thesis will then outline and detail the experimentation methods including the hardware for sensor pro- totypes and interfacing as well as testing setup for collision stimuli. Finally, the results from experimentation methods will be presented including analysis of sensor response to a large dynamic range of collision stimuli. Analysis of results is included to relate sensor response to modeling performed in the design chapter of the thesis. Following the results, the conclusion will highlight the work performed including the contribution of the design and analysis as 2 well as present future work associated with the sensor design. The conclusion will also cover the sensor advantages and contribution to the eld will be highlighted. 1.2 Contribution The sensor design allows for complex planar and non-planar applications. The sampled results show the following characteristics: high dynamic sensing range for collisions simple construction for ease of application to existing and future systems does not require rigid mountings or free space rigidity of surface is not a primary factor in response dynamics tactile shape and size is limited only by manufacturing technologies current technology allows for easily producible collision sensors based on the design and available parts The thesis work speci cally contributes to state of the art work by: Design and modeling for pseudo-membrane collision detection sensor. Provides for large area capabilities and robust applicability in contrast to membrane based approach [1], exible PCB with pressure approach [2], and MEMS based ap- proaches [4]. Achieves a high dynamic sensing range in contrast to tactile perception [5], limited sensing range [6], and object imaging applications [7]. Additional testing to support work published previously by the author [8]. 3 Chapter 2 Background In this chapter, the background information needed for later design understandings will be outlined. First, some initial information on the piezoelectric phenomenon used for sensing is presented and then the speci c material and phenomena utilized in the sensor construction is shown. Next, the current work being done for collision sensing and tactition is detailed to provide an accurate understanding of past and state of the art work in the eld. The classi cations and hazards associated with collision for robotics applications are then presented with speci c reference to international working standards and safety studies. Then, the motivations and applications of the sensor presented in the thesis are outlined with example robotic systems as well as general collision sensation. Finally, the coordinate system utilized in the thesis is presented to the reader. 2.1 Piezoelectricity and Ferroelectrics The word "piezoelectricity" comes from the Greek piezein, meaning to squeeze or press; thus, piezoelectricity is electricity related to squeezing or pressure. The phenomena of piezo- electricity was discovered and documented in 1880 by the Curie brothers through experimen- tation involving application of pressure to plates cut in a direction such that the production of an electric charge was observed [9]. Further quantitative analysis showed the piezoelectric e ect to yield charge proportional to the applied pressure to the crystals [9]. This direct piezoelectric e ect was also shown to work in reverse, allowing elements to be actuated with electricity. The piezoelectric matrix of constants involving the shear and normal vectors of stress and actuation was studied and detailed using maple, spruce and ash wood samples by Fukada in 1955 [10]. The majority of work with piezoelectricity involves the ferroelectric 4 property of polymers and ceramics. Ferroelectrics are polar materials possessing at least two equilibrium orientations of the spontaneous polarization vector eld of electric dipole moments of the material in the absence of external electric elds. The polarization vector may be switched between those orientations by application of an external electric eld [11]. In ferroelectric polymers, residual polarization due to the orientation of dipoles is stabilized and contributes to the pyro- and piezoelectric activities [12]. 2.2 Polyvinylidene Fluoride Polyvinylidene uoride is an inert high impedance polymer which can be formed into sheets as well as other easily manipulable forms and belongs to the ferroelectric polymer family. The chemical structure of PVDF molecules is given by (CH2-CF2). CF2 dipoles are aligned normal to the surface of the lm after poling and form a residual polarization [12]. The tensile piezoelectricity in stretched and poled lms of polyvinylidene uoride (PVDF) was rst demonstrated and documented by Kawai in 1969 [13]. This discovery triggered widely spread investigations on the pyro-, piezo-, and ferro-electricity of PVDF as well as its copolymers, nylons, and other polymers for many years [12]. Kawai [13] produced the piezoelectric e ect in PVDF lms by stretching the lms at high temperature, 100 C-150 C, and applying a static electrical eld to orient the dipoles. Sensing work with polyvinylidene uoride (PVDF) lm-based sensors includes tactile applications related to robotic skin for nger-tips [6], large area coverage [2], stress sensing for shock-wave measurements [14], de ec- tion sensing [15], object identi cation [7], and power harvesting [16] to name a few. Recent work with manufacturing has increased viability of PVDF as a exible and adaptable sensor solution for complex surfaces through MEMS-based fabrications [17] as well as the use of organic transistors to create a highly sensitive pressure sensor [18]. The work by Sekitani et al. [18] shows promising results utilizing organic transistors to boost signal output from PVDF sensors by integrating pre-amp stages to the signal into the exible polymer. The 5 reduced need for complex ampli er stages as well as signal loss concerns greatly increase the viability of exible PVDF-based sensors. 2.3 Collision Sensing Methods State-of-the-art methods for collision detection in robotics cover a wide variety of meth- ods and technologies, but for this thesis will be classi ed into subcategories related to robotics and environment interactions as well as human interfacing for control and input. The rst category uses knowledge of the statics and dynamics of the system, along with the current inertial or dynamic measurement of the movements, to detect collision. A second cate- gory of collision sensing, and the primary concern of this thesis, is the sensor design and implementation for collision detection as well as other tactition implementations. The col- lision technologies presented provide a point of reference to the reader for determining and understanding the bene ts of the sensor design and construction presented in Chapter 3. 2.3.1 Estimation and Control The estimation of collision using dynamic or inertial sensors inside the arm or system frequently incorporates control algorithms with the estimation. Xia et al. [19] utilizes joint torque sensors on a exible joint manipulator to estimate and implement control law for collision detection and planning. The collision detection presented is primarily focused on the manipulator or actuator of the robotic arm and precise control of the interaction with objects in an unstructured environment [19]. Je et al. [20] demonstrates collision detection utilizing an observer method which eliminates the need for external torque sensors. The work is again focused on the manipulator end of the robotic arm and primarily with grasp strength and intentional object interactions. The presented results show validity of the observer method using power supplied to the motor, measured by current consumption, related to the expected control input [20]. The main drawback of the estimation and control approach is the required knowledge of the system limiting the applicability; speci cally, the work using 6 torque sensors [19] requires extensive energy and elasticity information of the manipulator while the observer method [20] eliminates complex calculations, but necessitates knowledge of control input for the system. Work performed by Haddadin et al. utilized sensor-based collision detection to limit the force imparted by helper robots in a human-robot interaction environment [3]. Haddadin et al. successfully implemented a collision detection and reaction method which is capable of limiting force imparted by the robot to below 150N [3]. 2.3.2 Collision Sensors The sensor category of collision sensing is further decomposed to high delity sensors intended for complex tactition and large area and exible arrays for more complete coverage. Extensive work has been carried out in the eld of high resolution tactition using piezoelec- tric transducers. Kolesar and Dyson utilized a polymer lm array in 1995 to implement a tactile object imaging procedure for recognition of shapes including sharp edges, circle, and slotted screw. [7]. The work integrates a 64 sensor piezoelectric based array into an inte- grated circuit (IC) and performs pre-charge voltage bias to condition the load response. The experimentation yielded accurate object recognition; however, the tactile sensing elements have a low operating range of 0:00 N to 1:35 N [7]. Additional tactition research performed by Fujimoto et al., explored detection of slip with an aim towards implementation of arti- cial nger skin in robotics applications [6]. The results showed accurate detection of slip for static friction determination as well as the viability for the tactile sensing element to be incorporated in a robotic skin application; however, the sensor processing required complex arti cial neural network construction and calibration for the logic which limits the appli- cability of the design. Yamamoto et al. [5] utilized tactile elements created from PVDF to transmit surface texture utilizing a DSP. The work synchronized a tactile presentation display and tactile element to relay information of texture to a user to tactile telepresen- tation [5]. The primary sensor focus for the work was high delity of texture recognition and the sensors utilized were highly sensitive and specialized{limiting the dynamic range. 7 Extensive work by Seminara et al. to characterize the applicability of PVDF polymer lms for robotics applications showed promise for tactile integration and sensing [21]. The work?s primary concern was utilizing PVDF in a pressure-based sensing application to form exible tactile arrays for robotic skin, and the study showed promising sensing ability [21]. Further work in 2012 by Seminara et al. [2] showed the resulting polymer transducer array created from the design. The tactile array integrated PVDF transducers with exible printed circuit boards(PCBs) to build large area exible sensing systems. The study showed an acceptable level of sensation; however, the design requirements of exible PCBs and limited tactile size caused issues for large complex systems. 2.4 Hazard Classi cation and Concerns The primary concern for collision sensing and avoidance is elimination of risk not only to the user and environment but also the robotic system being integrated in dynamic coopera- tive environments. Collision sensing is utilized to satisfy safety requirements of power or force limiting for a robotic system in a cooperative environment [22]. The ISO standard 10218-1 [22] requires robotic systems to stop when a human is in the collaborative workspace of the system. The workspace of the robot for industrial applications involves a wide cordoned-o safety area; however, the workspace for non-structured environments is often de ned to be the mechanical area of the robotic system, including the space contained within the covers and collision sensors of a robotic arm. Examples of non-structured environments include the following: operating rooms, human helper robots in the home, medical imaging applications, and cooperative assembly lines. Collision sensations for standard satisfaction requires th sys- tem to limit overrun distance when collision occurs; therefore, detection of collision is more important than speci c force measurement. The International Standards Organization?s list of potential hazard origins for a robotic systems includes, but is not limited to, the following [23]: 8 movements of any part of the robot arm(including back), end-e ector mobile parts of robot cell rotational motion of any robot axes materials and products falling or ejection between robot arm and any xed object between end-e ector and any xed object unintended movement of machines or robot cell parts during handling operations The standard for robotic systems in an industrial environment goes on to list the following consequences or damage scenarios for the previous described hazards [23]: crushing shearing cutting or severing entanglement drawing-in or trapping impact stabbing or puncture friction, abrasion Major safety concerns exist for robotic cooperation in non-static human oriented environ- ments such as operating rooms, work places, emergency rooms, and medical imaging facil- ities. The majority of mechanical risks associated with robotics involve unintended touch or collision with objects and humans which creates the potential for harm and damage not only to the human or object but also to the robotic system; consequently, accurate collision detection to alleviate the risk is necessary for increased human and robotics cooperation. 9 2.5 Motivation and Application The sensor design and modeling presented in this thesis is primarily focused on elim- ination of sensor complexity present in state-of-the-art tactition sensors. Ease of use and robust applicability is of more concern for commercial safety systems than highly accurate tactition. With the increased use of complex robotics arms for human aid, simple systems for collision detection are necessary to insure safety of the operator and environment. The robotic system presented in Figure 2.1 shows an autonomous robotic imaging arm manufac- tured by the Siemens AG corporation. The system meets current safety standards through a mechanical switch and bumper based collision detection approach. Figure 2.1: Typical application for sensor design in non-structured environment. The shaded green area on the robotic arm represents desired collision detection areas. 10 Because of high dynamics associated with movement, the mechanical system has issues with false-positives, requires high levels of collision force to generate a detection, and has limited coverage area. The sensor design of the thesis would provide total system coverage of pinch points and other impact areas identi ed by the green shading in Figure 2.1 while also allowing for collision detection at much lower levels of force in order to reduce damage should collisions occur. 2.6 Coordinate System {+z, 3, thickness} {+y, 2, leng th} {+x, 1, width} Figure 2.2: Diagram showing Coordinate System used in Thesis The Cartesian coordinate system in this thesis uses the following equivalent relations interchangeably, fx;y;zg = f1;2;3g = fwidth;length;thicknessg. The diagram represent- ing the system for the sensor is shown in Figure 2.2. The positive z-axis is normal to the surface of the lm and extending away from the lm. The positivex-axis is extending away from the lm as shown while the positive y-axis is extending away from the lm and aligns 11 with the length direction of the lm. For notational simplicity, the vectors in-line with the axis will use a single notation reducing xx!x, yy!y, and zz!z. 12 Chapter 3 Design and Modeling of the Collision Sensor The primary focus of this chapter is the methodology, materials, and functions of the sensor design. By rst covering the merits of the materials and structure of the sensor, the reader can more clearly understand the motivations of the design. Next, the electrical e ects are explained and modeling of the sensor response to collision is derived to explicate the transduction of the physical model to measured electrical response. Finally, the electronics required for interfacing with and instrumentation of the sensor are detailed. 3.1 Piezoelectric Polymer PVDF is a piezoelectric and pyroelectric polymer commercially available in thin (<0.1 mm) sheets with uses including force sensors, accelerometer applications, high-frequency resonators, de ection sensing, and many more [15]. The piezoelectric PVDF lm is created from homopolymer PVDF sheets which are stretched, heated and simultaneously poled by application of a high voltage eld across the lm [24]. The stretching aligns the polymer chains of the PVDF and the high voltage orients the dipoles of the chains to create polar- ization in the lm [24]. The polling of the lm enables the polymer to generate charge when stressed by heat or physical stress because tensile stress in the lm causes the dipoles to ip, creating a charge gradient that generates an electrical displacement. The piezoelectric and pyroelectric e ects of the polymer do not signi cantly degrade over time (< 1% of original value) ensuring longterm reproducibility of sensations as long as the material is kept below approximately 90 C, depending on PVDF construction. At the Curie point, the poles of the polymer are randomly oriented eliminating the charge gradient [21]. PVDF polymer is uniquely suited to sensor applications due to commercial availability of poled and non-poled 13 PVDF sheets, readily available construction methods such as extrusion of lm, printing of electrodes and sputtering for more delity of measurements,, and the material properties of inertness, elasticity and durability. 3.2 Piezoelectric E ect The proposed sensor utilizes the piezoelectric e ect of PVDF thin lms to transduce a sensing response to collision with the system. The tactile element transduces experienced stress to an electrical displacement, D, which is the charge density of the lm surface, Q=A. The electrical displacement created has additive components of pyroelectric, piezoelectric, and dielectric e ects [11]: D = p T +djkXjk +"E (3.1) The pyroelectric charge is a function of the change in temperature ( T) times pyroelectric charge coe cient (p), the piezoelectric charge is a relation of stress applied in Cartesian direction (Xjk) with the corresponding piezoelectric charge coe cient (djk), and the charge related to electric dipole moment is calculated by electric eld (E) times the permittivity of the material ("). For collision sensing, the desire is for electrical displacement, D, to be a purely piezoelectric response, djkXjk. The piezoelectric response is isolated through design and ltering in the following manners. The electric eld, E, can be minimized by proper design of the charge ampli er sensor interface in order to eliminate build up of the eld. The charge ampli er is discussed in detail later in Subsection 3.7.1. The pyroelectric component, T, can be canceled because of phase orientation of the bilayer sensing element and common mode signal ltering. The elements are placed such that electrodes are reversed on the top and bottom layer resulting in the pyroelectric e ect being positive for the upper element and negative for the lower element. The response is out of phase by 180 degrees and the signal can then be ltered by rejecting the common mode of the combined signal. The stress response of the elements is a signed response dependent on the de ection of the lm 14 resulting in isolated piezoelectric response when the lms are de ected in the same direction. Therefore, the pyroelectric and dielectric e ects fall away reducing (3.1) to the desired purely piezoelectric displacement in (3.2). D = djkXjk = d31X31 +d32X32 +d33X33 (3.2) The stress vector, Xjk, in Equation (3.2) represent tensile stress in length, width, and thick- ness directions respectively. Due to high compressibility of the substrate relative to the PVDF lm, strain related to compression, X33 is approximately 0. The piezoelectric con- stants corresponding to tensile stress in the width and length, d31 and d32 respectively in Cartesian coordinate representation1, are equal. Therefore, the electrical displacement of the sensor is proportional to the total transverse and longitudinal stress in the lm created by the collision reducing Equation (3.2) further to the reduced representation of the sensor electrical displacement in (3.3). D = d31(X31 +X32) (3.3) The applicable material properties of PVDF lm are shown in Table 3.1 and provided by the lm manufacturer, Measurement Specialties. Table 3.1: Material Properties of PVDF Film E Young?s Modulus 2 4 109N=m2 d31 Transverse Coe cient 23 10 12C=m2N=m2 d33 Compressive Coe cient 33 10 12C=m2N=m2 p Pyroelectric Coe cient 30 10 6 Cm2K 1Recall the Cartesian coordinate system in this thesis uses the following equivalent relations interchange- ably,fx;y;zg=f1;2;3g=fwidth;length;thicknessg. The diagram shown previously in Figure 2.2 describes the system in reference to a material sensor. 15 3.3 Membrane Construction Collison Element 1 Element 2 Figure 3.1: Membrane Sensor Construction Diagram The membrane based construction shown in Figure 3.1 transduces collision to an elec- trical displacement through stress application to a rigidly mounted membrane. The sensor construction generates stress in tension because of de ection of the membrane from applied force. The equations for modeling the sensation of force collisions are easily derived because the stress, shown in Equation (3.4), is purely a function of the force over cross-sectional area of the lm. S = FA c (3.4) The pseudo membrane approach involves stress in length and width orientations of the ele- ment. The membrane experiences stress in primarily the length direction because increased cross sectional area in the width direction limits the response for membranes with length much greater than the width. Recall that the piezoelectric e ect is a function of the piezoelectric constants and applied stress; therefore, the generated electrical displacement, shown in Equation (3.5), for the membrane approach is congruent to the pseudo-membrane results, discussed later in Section 16 3.6 and shown in Equation (3.19). D = d31S (3.5) Furthermore, the charge (Q) created can be derived by multiplying the electrical displace- ment, which is a charge density, by the area of the element (A). Q = d31 FA c A (3.6) System modeling is greatly simpli ed for the membrane approach. However, the membrane construction method su ers in an application sense because it requires free space for di- aphragm de ection and rigid mountings in order to achieve dynamic sensing range. These requirements necessitate specialized construction on an application by application basis, lim- itation in sensor curvature and reduction of thickness, and reduced durability because the sensor relies on the elastic nature of the lm to return to steady state. Large repetitive force application can lead to degradation in performance over time. Therefore, in this thesis the novel approach will utilize a pseudo-membrane construction to eliminate construction requirements while maintaining similar sensing method to a membrane approach. 3.4 Pressure Construction The pressure-based construction, shown in Figure 3.2, relies on stress created from pressure applied to the sensor. The pressure is generated because the sensor is mounted to a su ciently rigid substrate or surface such that force applications cause compression of the lm elements. The stress(S), shown in Equation (3.7), is now the pressure generated by the collision or the force(F) of the collision over the contact area(Ao) of the object. S = FA o (3.7) 17 Collison Element 1 Element 2 Figure 3.2: Pressure Sensor Construction Diagram The presence of the contact area component means that smaller objects generate higher pressure for equal force, which can be problematic for large area sensors. The electrical displacement(D), shown in Equation (3.8), now depends on the pressure piezoelectric con- stant (d33). D = d33S (3.8) The charge can be derived in a similar manner to the membrane approach, and the problem related to object size can be eliminated by making the sensor area su ciently small such that the force is evenly applied over the area of the sensor resulting in the cancellation of the area terms. The canceling of the areas(A&Ao) yields a charge result(Q) directly proportional to the force(F) input, shown in Equation (3.9). Q = d33 FA o A = d33F (3.9) The construction eliminates the need for rigid mountings and free space of the membrane; however, the sensor output of strain due to pressure is much lower yielding more complex and specialized electronics. The pressure-based construction also requires a very rigid surface and/or substrate to generate pressure transduction. 18 3.5 Sensor Structure Substrate Collision Element 1 Element 2 Figure 3.3: Pseudo-Membrane Sensor Construction Diagram The collision sensor developed in the thesis is constructed from two PVDF lm ele- ments oriented with poles out of phase adhered to a exible elastic compressible substrate. The trilayer sensor is attached to the targeted surface, shown in Figure 3.3. The construc- tion yields a pseudo-membrane operation state in contrast to the traditional membrane and pressure-based design included as Figure 3.1 and 3.2, respectively and discussed in the previous sections . The collision stimulus deforms the elastic substrate due to localized compression and creates a resulting mechanical strain on the PVDF lm elements that is mathematically related to the applied collision force in the next section. The elements act semi-membrane-like because the stress is an e ect of the de ection due to compression. The elastic substrate should be chosen or designed to maximize the linear stress strain response 19 and also to minimize total sensor size for manufacturing and application concerns. The tri- layer pseudo-membrane approach using PVDF and an elastic substrate are capable of large area coverage because large PVDF sensing elements are easily constructed, the elastic sub- strate can be made of polyurethane foams and other commercially available compressible materials, and the sensor is not limited to planar surfaces because the pseudo-membrane approach creates stress from localized substrate compression and not lm de ection as in a membrane approach. Recall the membrane based construction in Figure 3.1 required a certain amount of free space for de ection as well as rigid mountings. Also, recall the pres- sure construction in Figure 3.2 eliminated these issues but yields lower response and requires increased surface rigidity. The pseudo-membrane construction in Figure 3.3 eliminates the specialized construction requirements of rigid mountings for the membrane, free space for de ection, and rigidity of the surface a ecting dynamic range. The dynamic range of the sensor and response are now controlled by the substrate material properties related to the thickness. However, the surface should be more rigid than the compressible substrate such that compression occurs, and the substrate should not be so hard or incompressible that the lm elements begin to act in a pressure sensing capacity rather than membrane-like one due to de ection. 3.6 Response Modeling For modeling purposes, the area of concern is restricted to the local frame of the collision. To maintain linearity of the response we will assume that the collisions will stay within the approximately linear region of the stress-strain response curve [25]. Also, the substrate will be treated as a continuum. The lm stress from (3.2) is equal to the stress of the surface of the substrate, assuming a perfect adhesive bond and negligible e ects of lm sensor on substrate stress characteristics. For the reader?s bene t, the material properties and modulus 20 utilized in the following response modeling are detailed and de ned in Appendix A. = 2 66 66 4 Te1 Te2 Te3 3 77 77 5 = 2 66 66 4 11 12 13 21 22 23 31 32 33 3 77 77 5 = 2 66 66 4 x xy xz yx y yz xz yz z 3 77 77 5 (3.10) By de ning the Cauchy stress tensor ( ) of the substrate, (3.10), in terms of the normal and shear stresses, f x; y; zg and f xy; xz; yzg, Xjk in (3.2) can be replaced by the stress vector, Te3, corresponding to the material surface resulting in Equation (3.11). D = d31( xz + yz) (3.11) Equation (3.11) shows the electrical displacement, D, as the piezoelectric constant corre- sponding to orthogonal stress multiplied by the combination of the shear stress in the length and width direction, xz and yz, respectively. From Equation (3.11), the stress contribut- ing to the piezoelectric e ect of the lm is the orthogonal shear stress experienced by the substrate at the collision point. Therefore, the sensor dynamic range is dependent on the shear and normal stress characteristics of the chosen substrate. Using the shear modulus of elasticity (G) to relate shear strain to shear stress and Young?s Modulus of the substrate (E) to relate normal strain to normal stress (3.12), G = xz xz E = z z (3.12) along with the geometric representation of strain, (3.13), ij = 12( ij o + ji o ) (3.13) 21 and Cauchy?s strain tensor ( ) (3.14), = 2 66 66 4 11 12 13 21 22 23 31 32 33 3 77 77 5 = 2 66 66 4 x xy2 xz2 yx 2 y yz 2 xz 2 zy 2 z 3 77 77 5 (3.14) Assuming the substrate is under compression locally where x = y = 0, z 6= 0, and xo;yo are known static quantities, the shear stress is transformed to a compressive stress, using the relations (3.12) through (3.14). First, the shear stress of the lm, , is transformed to the engineering shear strain, using the shear modulus of elasticity relation. The result- ing engineering shear strains are equivalent to shear strains of the material because of the properties of the Engineering Strain Tensor outlined in Appendix A.2. xz + yz = G( xz + yz) = G(2 xz + 2 yz) (3.15) Next, the shear strains are converted to geometric representations, and the fractions from Equation (3.13) cancel because of relations in the strain tensor (3.14). The resulting Equation (3.16) is now a function of the change in thickness, compression, of the material and the original dimensions{length and width. xz + yz = G( zx o + zy o ) (3.16) By combining the fractions and adding a term of the original thickness, the stress of the surface of the lm is transformed to a proportionate function of the compressive strain of the material, z, shown in Equation (3.17). xz + yz = Gzo(yo +xo)x oyo ( z) (3.17) 22 Finally, the compressive strain of the material is transformed from Equation (3.17) to com- pressive stress in Equation (3.18) such that the stress of the lm, S, becomes a proportionate function of the compressive stress of the material caused by the collision of the object. xz + yz = GEzo(yo +xo)x oyo ( z) = S (3.18) In this case, compressive stress, z, is a monotonically increasing and directly proportionate function of the force of the collision normal to the sensor where force towards the sensor produces a positive response. Speci cally, higher levels of collision force should yields higher levels of compressive stress for the same object and lower levels of collision force resulting in lower levels of stress in the locally compressible area of the object in collision. By substituting the derived stress of the lm sensor, S, into the piezoelectric electrical displacement formula from Equation (3.11) , the resulting sensor response is expressed in Equation (3.19) D = d31( xz + yz) = d31GEzo(yo +xo)x oyo ( z) = d31(S) (3.19) Therefore for a measured stress S in Equation (3.19) and the electrical displacement of the pseudo-membrane sensor should also be a monotonically increasing function of the force. Modeling of strain was primarily accomplished with reference to [25]. 3.7 Instrumentation 3.7.1 Ampli er Electronics The electronics interface for the PVDF lm elements requires high signal gain, low out- put impedance, high input impedance, low time constant to capture 1Hz collisions, and a minimization of the electric eld e ect of the sensor. A charge ampli er is used to minimize e ects of sensor and line capacitance by minimizing input impedance, to minimize elec- tric eld by grounding sensor electrode, and because the sensing elements act as a current 23 ? + + V cc ? V cc C R V out C s Is Sensor Mo del V in Figure 3.4: Charge Ampli er Circuit Design source. The circuit, shown in Figure 3.4, acts as a single pole high-pass lter with a bleed resistor added in parallel to create a low enough cuto frequency to properly detect physical interaction in the 1Hz - 1000Hz range [21]. The transfer function is shown in (3.20): H(s) = Vout(s)V in(s) = sRCssRC + 1 (3.20) The system can be properly designed to yield a low enough corner frequency calculated from (3.21), fc = 12 RC (3.21) which gives the needed low-end frequency range. The signal is then low-pass ltered with a cut-o frequency to attenuate unwanted signals and conditioned for input to the analog to 24 digital converter (ADC). The charge ampli er allows for positive and negative voltage range of +/- Vcc to create the large dynamic range needed for collision detection. 3.7.2 Sampling and Detection Electronics The collision detection and sampling of the post ampli ed signal is performed in multiple ways. Threshold level collision detection is implemented utilizing a comparator and tunable reference voltage. The threshold voltage level is chosen based on signal response for desired collision force level. The comparator circuit works by outputting a \1" when the sensor signal exceeds the voltage threshold and then a \0" when the signal returns below. The circuitry is easily interfaced with solenoids or electronic interrupt hardware to create a stop function, satisfying industrial safety requirements [22]. Sub-threshold level collision detection is ac- complished digitally by feeding the sensor signal into ADC?s. The ADC inputs require signal conditioning to achieve full sensor dynamic range; speci cally, the signal from the ampli ers should be recti ed to the voltage range available to the ADC and then shifted such that all signal data is captured. The circuitry involves a passive voltage recti er is implemented with resistors and an active voltage shifting circuit using a unity gain non-inverting opera- tional ampli er and an o set reference voltage. The ADC converts the conditioned signal to digital values which can then be processed to detect collision events which occur below the threshold or sampled and sent back to a computer for data logging. For the experimentation and results detailed in the following chapters, the ADC and resulting digital values are used in the sampling format. 3.8 Design Attributes The sensor design allows for complex planar and non-planar applications. The pseudo- membrane allows for high dynamic sensing range and applications while eliminating the construction requirements of membrane and pressure sensors. Recall that the membrane sensor construction requires free space for diaphragm de ection as well as rigid structures 25 for mounting the sensor element. The pseudo-membrane eliminates the need for free space by utilizing a locally compressible substrate and the rigid mounting is also not needed because adherence to the substrate creates mounting for sensor stress during de ection. Recall the pressure based construction eliminated free space needs, but required a highly-rigid substrate or surface to generate acceptable dynamic range. The pseudo-membrane maintains the ease of use of pressure actuation, but eliminates the high rigidity requirement because the sensor is acting in a de ection mode. The sensor design di ers from state of the art tactile and collision sensing design in the following manner: high dynamic sensing range for collisions simple construction for ease of application to existing and future systems does not require rigid mountings or free space rigidity of surface is not a primary factor in response dynamics tactile shape and size is limited only by manufacturing technologies current technology allows for easily producible collision sensors based on the design and available parts The experimental methods and results presented in the following chapters will expound upon published work on the sensor design by more accurately characterizing the force of the object during impact and additional statistical analysis on previous results [8]. The sensor design provides a novel collision sensing solution for complex robotics and automated systems where safety and system coverage are of higher concern than high resolution tactile perception. 26 Chapter 4 Experimentation Methods and Prototype This chapter will outline the prototype sensor and electronics utilized in testing. Also, the experimentation methods used will be shown. The prototype sensor is constructed to represent a planar and non-planar application for the design. The electronics are detailed including the speci ed frequencies for ampli er design. The approximate force experimen- tation for collision simulation is detailed. Also, a more accurate force measurement method is presented utilizing high speed cameras and image correlation techniques. 4.1 Prototype Sensor Initial prototype sensors for testing were constructed from poled 28 m thick PVDF lm elements, each 171 mm by 19 mm (length and width) and a 0:5 inch polystyrene closed cell foam substrate. The polyurethane foam was chosen for availability and to allow for large amounts of compression at collision. The trilayer sensor was constructed by adhering the two lms, Element 1 and Element 2 from Figure 3.3, out of phase such that Element 1?s top electrode is positive and Element 2?s top electrode is negative, adhering the bilayer PVDF lm to the polystyrene foam substrate and then a xed to the sample robotic arm shield for collision testing. The prototype sensor is shown in Figure 4.1 with curved portions and edges representing non-planar applications and the at parts of the prototype representing planar applications. 4.2 Prototype Electronics Signal capture was performed using previously described ampli er circuit design inter- faced to 12-b analog to digital converters on an Atmel Xmega microcontroller using a bu er 27 Figure 4.1: Sensor Prototype used in testing showing planar (top of cover) and non-planar (rounded left end) sensor applications on example robotic arm shielding. Wires in picture connect sensor electrodes to instrumentation. and signal conditioning ampli er stage. The charge ampli er was designed with a 1:6 M bleed resistor and 100 nF charge accumulating capacitor yielding the following corner fre- quency: fc = 12 RC = 12 (100 nF)(1:6 M ) = 0:997 Hz (4.1) which gives the needed low-end frequency range. The signal is then low-pass ltered with a 1 kHz cut o frequency to attenuate unwanted signals. The charge ampli er allows for a positive and negative voltage range of 15 V to 15 V to allow for high gain and large dynamic range. Because the ADC operates from 0 V to 3 V, the signal output of the charge ampli er 28 is regulated down to the 1:5 V to 1:5 V range and then shifted up 1:5 V. The ADC uses a reference voltage of 1:5 V in di erential mode resulting in a digital output range of 2048 to +2048 corresponding to a VLSB of 7:32 mV. The ADC?s sampling frequency is 93:7 kHz which is more than 40 times the bandwidth of the analog input. Data was logged using serial communication with signals down-sampled to 10kHz. 4.3 Method 4.3.1 Collision Simulation and Approximate Force Collision stimuli for testing was generated by dropping an object of known weight and uniform contact area on the sensor from varied heights to produce controlled impact colli- sions. Average Force of the object during impact (F) is approximated using the conservation of energy, shown in Equation (4.2), and the work-energy principle, shown in Equation (4.3), where distance to slow down (d) is compression of the substrate, de ned as compressive strain multiplied by thickness. mgh = 12mv2 (4.2) Work = F d = 12mv2 = mgh (4.3) The work energy principle states that work is equal to the change in kinetic energy meaning that average impact force over the slow down distance is therefore equal to the original potential energy of the mass. The height(h) includes the slow down distance and height of object from the surface. An approximation of distance for the object to slow down is a 50% compression of the substrate resulting in a slowdown distance of 0:25 inches. The approximate method allows for comparison of collision responses which closely simulate real world events; however, the approximate average force output has some inherent inaccuracy because of nonuniform compression of the material, variations of accelerations due to the non-uniformity over the range of collision stimuli, and bounce during impact. The average force due to work-energy 29 principle assumes the object is stopped in impact and does not account for non-zero velocity which occurs during rebound; therefore, the average force during impact is likely higher, especially for larger collisions where excessive bounce occurs. The method does not provide exact force estimation{particularly for the upper dynamic range of the sensor; however, he experimental method provides a good reference of response from smaller and larger levels of collision force to characterize the dynamic range of the sensor design and substrates; Table 4.1 shows the collision stimuli and approximate average force generated from the impacts. Objects used in testing were a Craftsmen wrench in size 15 mm and 22 mm, for types 1 and 2 respectively. Table 4.1: Collision Stimuli Data Set Object Type Weight Object Height Average Force A I 353 g 1:5 cm 5 N B I 353 g 3 cm 10 N C I 353 g 5 cm 20 N D I 353 g 8 cm 30 N E I 353 g 10:2 cm 40 N F II 597 g 10 cm 60 N G II 597 g 13 cm 80 N H II 597 g 16:3 cm 100 N I II 597 g 32:6 cm 200 N J II 597 g 49 cm 300 N 4.3.2 Force Measurement For more accurate force representation of collision response, a secondary system is re- quired to measure the displacement, velocity, and acceleration of the objects at impact to characterize the impacts. In order to limit the e ects additional sensors would have on the physical dynamics of the model, object displacement, velocity, and acceleration measure- ment is performed using Digital Image Correlation (DIC) and a high speed camera. DIC is useful because of accuracy, computational e ciency, and elimination of the need to attach additional sensors [26]. The method utilizes a random spray or placement of particles as tracking points and correlates the movements of the points from image to image in order to 30 calculate the displacement from frame to frame. The measurement output for the system is the displacement of the object while in the frame of view of the camera. Acceleration is derived from the displacement by performing a line t and taking the second derivative of the resulting polynomial. The collision stimuli for this force measured approach are presented in Table 4.2. The set is intended to mimic the coverage of the previous data set, and the presented average forces cover the range for the approximated method shown in Table 4.1. Table 4.2: Collision Stimuli for Displacement Measurement Tests Data Set Weight Object Height Average Force A 500 g 5 cm 38:58 N B 500 g 10 cm 77:17 N C 500 g 15 cm 115:75 N D 500 g 20 cm 154:33 N E 500 g 35 cm 270:08 N F 500 g 50 cm 385:83 N The testing setup utilizes a precision machined cylindrical weight with a rounded end in order to prevent sensor damage. The weight is machined to within 0:1 g of 500 g to simplify force calculations. The test setup is pictured in Figure 4.2. The testing methods are designed to mimic real world impact collisions spanning the desired operating range of the sensor. The approximate method provides for e cient testing to con rm the planar and non-planar applications of the sensor. The presented force mea- surement method provides additional impact information of the objects during collision to support the results of the approximate method as well as characterize the impact character- istics of the foam. The prototype sensor represents the planar and non-planar applications for testing. 31 Figure 4.2: Experimental setup using High-speed(10000 fps) camera and robotic arm to create and measure consistent collision stimuli. The robotic arm is located over the sensor prototype which is clamped to the table to provide rigidity and is shielded from the heat of the lights by the illuminated wall in the front. The camera is behind the two 1000 W lights used to increase the contrast ratio of the falling object. The entire setup is contained in a light box environment to eliminate external interference. 32 Chapter 5 Results This chapter presents and analyzes the results of the experimentation methods presented in the previous chapter. The results for the approximate force method show the sensor response to the range of collision stimuli described in Subsection 4.2.1 and shown in Table 4.1. For this result set, a planar sensor prototype and non-planar sensor prototype are tested, and the resulting responses are compared. Experimentation results from the force measurement method are presented for a planar case. The data set presented covers a similar collision range to the approximate method and has been previously provided in Table 4.2. 5.1 Average Force Experimentation Method Recall from Subsection 4.3.1, the experimentation utilizing average force is carried out by generating multiple collision stimuli from the heights speci ed in Table 4.1. The multiple collision tests are performed on both planar and non-planar sensor prototypes, and the resulting signals are captured. The mean response of both sensor prototypes is taken, and the peak of the mean response is utilized in comparison. Statistical analysis is also performed on the resulting peak measurements to characterize the sensor response to varying levels of collision force for both sensor prototypes. 5.1.1 Planar Sensor Prototype The sampled mean results, presented in Figure 5.1, show the wide dynamic sensor range, from 5 N to 300 N, and consistent response to collision. The mean response for each collision force level is shown on the graph. The sensor response is the value of S derived from the 33 measured electrical displacement, previously shown in Equation (3.19). Recall the planar sensor prototype is the at area of the example cover previously shown in Figure 4.1. 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0 2 4 6 8 10 12 14 16 18 20 x 109 S (Pa ? 10 9 or GPa) Time(s) Collision Responses 300N 200N 100N 80N 60N 40N 30N 20N 10N 5N Figure 5.1: Mean captured results for wide dynamic range of collision stimuli for planar application of sensor. The relation of applied collision stimuli to measured stress peaks resulted in Figure 5.2 showing that the sensor response is not totally linear. The relation resembles the engineer- ing stress strain curves of polyurethane foam under impact, which reinforces the previous assertion that stress measured by the sensor is related to the localized compressive strain at the impact point [27]. 34 0 50 100 150 200 250 300 0 2 4 6 8 10 12 14 16 18 20 x 109 S (Pa ? 10 9 or GPa) Collision Force(N) Response vs Force Figure 5.2: The relation of measured collision to approximate force of object collision for planar application shows an initial linear trend of high sensitivity but reaches a non-linear region and eventual plateau of decreasing strain experience and reduced sensitivity for force greater than 100N. 35 5.1.2 Non-planar Sensor Prototype The results in Figure 5.3 show that the measured response from collision for a non-planar application strongly correlate with the results from planar application. Recall the non-planar sensor prototype is the curved end of the example cover previously shown in Figure 4.1. Some deviation is expected, but the sensor provides detection over the desired dynamic range for the non-planar application. The consistency between planar and non-planar experiments demonstrates the robust application properties of the sensor. The relation of measured system response to approximate collision force, shown in Figure 5.4, is also very similar to the graph from the planar testing exhibiting an initial linear region and then plateau. 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0 2 4 6 8 10 12 14 16 18 20 x 109 S (Pa ? 10 9 or GPa) Time(s) Collision Responses 300N 200N 100N 80N 60N 40N 30N 20N 10N 5N Figure 5.3: Mean captured results for wide dynamic range of collision stimuli. 36 0 50 100 150 200 250 300 0 2 4 6 8 10 12 14 16 18 20 x 109 S (Pa ? 10 9 or GPa) Collision Force(N) Response vs Force Figure 5.4: The relation of measured collision to force of object collision for sensor mounted on a non-planar surface shows an initial linear trend of high sensitivity but reaches a non- linear region and eventual plateau of decreasing strain experience and reduced sensitivity for force greater than 100N, very similar to the response graph of the planar application. 37 5.1.3 Low Force Dynamic Response Figure 5.5 shows that there is little delay or di erence in stress measured by the upper and lower sensors; however, some di erence can occur because of de-lamination of the sensor elements from each other after repeated impacts with inadequate construction. Figure 5.5 also shows the uniformity of sensor response for the top and bottom sensor elements. The level of the digital response for the low impact should make it clear to the reader that low levels of collision are highly detectable by the sensor. The 5 N collision force of the impact generated a digitally quantized value in excess of 200, which represents the amplitude of the analog sensor signal. The quantized voltage has a least signi cant bit equivalent value of 1 = 7:32 mV. 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04?100 ?50 0 50 100 150 200 250 Time(s) Digital Response(V LSB ) Measured Digital Response of 5N Collision Top Element Bottom Element Figure 5.5: Mean Digital Response results of Element 1 and Element 2 for 5N collisions. 38 5.1.4 Total Results Comparison and Statistical Analysis The complete data set values are compiled and shown in Table 5.1. The mean peak of the sensor response for planar and non-planar sensor prototypes have similar magnitude for each object drop height. The standard deviation for the planar sensor measurements exhibits the strong correlation and response density such that results between di erent force levels can be accurately distinguished. The higher levels of deviation in non-planar sensor response is likely resultant of some non-normal force during impact with the curved sensor application. The average standard deviation for the range of collisions levels for planar sensor prototype is 0:4079 GPa. The non-planar results, which exhibit the previously discussed deviations, have an average standard deviation of 0:7098 GPa. The compiled results show the preciseness of the sensor response for planar and non-planar applications. However, additional force measurements are necessary to accurately model the object impacts and sensor response, given in the following section. Table 5.1: Total result comparison of planar and non-planar sensors with mean, , and standard deviation, s, for each force stimuli set. Data Height Force p (GPa) sp (GPa) np (GPa) snp (GPa). A 1:5 cm 5 N 4.576 0.267 3.466 0.441 B 3 cm 10 N 5.695 0.265 4.557 0.794 C 5 cm 20 N 7.664 0.362 6.427 0.656 D 8 cm 30 N 8.864 0.422 7.339 0.676 E 10:2 cm 40 N 10.078 0.395 8.970 0.762 F 10 cm 60 N 12.470 0.413 11.686 1.010 G 13 cm 80 N 13.792 0.392 14.157 0.667 H 16:3 cm 100 N 14.392 0.556 15.284 0.669 I 32:6 cm 200 N 17.663 0.520 17.400 0.864 J 49 cm 300 N 18.814 0.487 19.923 0.559 39 5.2 Force Characterized Collision Response Method In this section, the falling object is tracked utilizing a 10000 frame-per-second camera, and the displacement is derived from the videos utilizing the image processing method de- scribed in Subsection 4.3.2. The displacement measurement for the object falling, impacting the sensor, and then rebounding is shown in Figure 5.6. Figure 5.6 shows measurements for a 5 cm drop and the di erent phases of the object movement are highlighted. 0 0.05 0.1 0.15 0.2 0.25?0.06 ?0.05 ?0.04 ?0.03 ?0.02 ?0.01 0 Displacement Measurements for 5CM Drop Displacement(m) Time(s) Measured Displacment Free Fall Impact Figure 5.6: Measured displacement for 5 cm drop. The displacement measurements corre- sponding the object in free fall and impact are highlighted in red and blue respectively. 40 The free fall portion, highlighted in red, is utilized to con rm measurement validity by estimating acceleration due to gravity, which is detailed in Subsection 5.2.1. The portion of importance for collision force measurement{the impact period{is highlighted in blue. The displacement measurement during this impact region is used to calculate the accelerations of the object during impact, and using the mass of the object, the resulting peak impact force is derived. The method for force calculations from displacement measurements is detailed in Subsection 5.2.2. 5.2.1 Gravity Con rmation In order to verify the method of displacement and force measurement, the object move- ment is captured in free fall prior to impact. Using the free fall displacement window, the acceleration due to gravity is achieved by a second order line t to the displacement mea- surements and sample validity is determined. Over the range of collision stimuli speci ed in Table 4.2, the estimated gravity during free fall has a root-mean-square error of 0:275 79 m/s2 and mean value of 9:7815 m/s2 The gravity measurements are shown in Figure 5.7. 5.2.2 Force Measurement The force of the object during collision is calculated from the measured displacement during the impact of the object. The line t of the displacement data is utilized in a similar fashion to the gravity con rmation method; however, the polynomial line t is of higher order and the primary concern is the peak response. The relation of measured peak response for the data sets to the height of the object dropped is shown in Figure 5.8. The measured force response to height follows a monotonically increasing trend. The force levels shown in Figure 5.8 do not achieve the levels of strain to clearly show the plateau of realizable force which will occur at higher levels of impact. The results show that peak force impact is higher than estimated average impact from Table 4.2, as expected, because object rebound is not compensated for in the average impact force derivations in Equation 4.3. 41 0 5 10 15 20 25?10.4 ?10.2 ?10 ?9.8 ?9.6 ?9.4 ?9.2 ?9 Gravity Measurements RMSE= 0.27579 Acceleration(m/s 2 ) Sample( #) Measured Gravity Truth Gravity Figure 5.7: Gravity measurements made to con rm validity of displacement measurements with a root-mean-square error of 0:275 79 m/s2 and mean value of 9:7815 m/s2 . 42 0 50 100 150 200 250 300 350 400 4505 10 15 20 25 30 35 40 45 50 Drop Height (cm) Impact Force (N) Mean Peak Impact Force vs Drop Height Figure 5.8: Mean of measured peak impact force for drop height of object during collision stimulus. 43 5.2.3 Sensor Response The sensor response for the object impacts for each data set were captured in order to characterize the response to measured peak impact force. The sensor response resulting from the 5 cm object drop are shown in Figure 5.9. The measured values is the S component previously discussed in Equation (3.19). The sensor response shows strong uniformity and correlation for multiple impacts and a close grouping of peak response level. The sensor responses for 10 cm, 15 cm, and 20 cm object drop heights are shown in Figure 5.10 , 5.11, and 5.12 respectively. 0 0.005 0.01 0.015 0.02 0.025?0.5 0 0.5 1 1.5 2 2.5 3 3.5x 10 9 S (Pa ? 10 9 or GPa) Time(s) Collision Response for 5 CM Drop N905172 N905173 N905174 Figure 5.9: Measured sensor response for 5 cm drop with responses labeled by data run notation of [N9 Date Run#]. 44 0 0.005 0.01 0.015 0.02 0.025?1 0 1 2 3 4 5 x 10 9 S (Pa ? 10 9 or GPa) Time(s) Collision Response for 10 CM Drop N905175 N905176 N905177 Figure 5.10: Sensor response for 10 cm drop with responses labeled by data run notation of [N9 Date Run#]. The sensor response for heights ranging 5 cm to 20 cm all show similar uniformity and peak point grouping. The sensor responses in Figure 5.11 and 5.12 exhibit exceptional uniformity and tight grouping for peak sensor response. The deviation present in other responses is likely a result of object impact not being perfectly centered on the lm sensor. 45 0 0.005 0.01 0.015 0.02 0.025?1 0 1 2 3 4 5 6 7 x 10 9 Time(s) S (Pa ? 10 9 or GPa) Collision Response for 15 CM Drops N905178 N905179 N9051710 N9051711 Figure 5.11: Sensor response for 15 cm drop with responses labeled by data run notation of [N9 Date Run#]. 46 0 0.005 0.01 0.015 0.02 0.025?1 0 1 2 3 4 5 6 7 8 x 10 9 S (Pa ? 10 9 or GPa) Time(s) Collision Response for 20 CM Drop N9051712 N9051713 N9051714 Figure 5.12: Sensor Response for 20 cm drop with responses labeled by data run notation of [N9 Date Run#]. 47 The measured response corresponding to object impacts created from 35 cm and 50 cm heights exhibit the previously discussed deviation from object impacts not squarely colliding with the sensor. The low response in Figure 5.13 can be excluded from response mean calculations because it is clearly an outlier resulting from the object not cleanly impacting the sensor. The high dynamics of impacts for the object at the higher drop heights and force levels create deviation in the sensor response. 0 0.005 0.01 0.015 0.02 0.025?1 0 1 2 3 4 5 6 7 8 9 x 10 9 S (Pa ? 10 9 or GPa) Time(s) Collision Response for 35 CM Drop N905211 N905212 N905213 Figure 5.13: Sensor response for 35 cm drop with responses labeled by data run notation of [N9 Date Run#]. 48 The responses for 50 cm drops, shown in Figure 5.14, show extreme deviation such that an accurate mean cannot be determined. However, the measured peak force for the higher levels of dropping does not exhibit the same amount of deviation. It is likely that variation in impact point between the object and sensor are the root cause of deviation for similar levels of measured force. The force measurements do not deviate because the foam substrate generates the same amount of response both on and o of the sensing tactile. 0 0.005 0.01 0.015 0.02 0.025?1 0 1 2 3 4 5 6 7 8 9 x 10 9 S (Pa ? 10 9 or GPa) Time(s) Collision Response for 50 CM Drop N905214 N905215 N905216 N905217 N905218 Figure 5.14: Sensor response for 50 cm drop with responses labeled by data run notation of [N9 Date Run#]. The data runs show extreme deviation because of object impacts partially impacting the tactile. 49 The measured sensor response for the range of collision detections, shown in Figure 5.15, shows the strong correlation of some sensor response levels while others exhibit more deviation. Unlike the average force experimentation method where large numbers of collisions were created (more than 30) for each level, the complexity of camera based displacement measurement and processing limited the number of achievable collisions. The low number of data points causes problems for sensor response measurement when the object does not squarely strike the sensor, as previously discussed. 0 50 100 150 200 250 300 350 400 4502 3 4 5 6 7 8 9 x 10 9 S (Pa ? 10 9 or GPa) Peak Impact Force (N) Peak Collision Responses vs Peak Force 5cm 10cm 15cm 20cm 35cm 50cm Mean Figure 5.15: Sensor response vs measured impact force for total data run including mean value line. 50 The trend of sensor response versus measured peak impact force strongly supports the previous results achieved using the approximate force method. The measured force detection of 49:315 N for the 5 cm shows the sensor is capable of low levels of collision detection down to 49:315 N. The detection at 413:778 N, while slightly inaccurate, clearly demonstrates the wide dynamic range of the sensor design. The results of collision response and force measurement are numerically represented in Table 5.2. The calculated mean of each collision level as well as sensor response are shown. The range of collision measurement and response is utilized in lieu of standard deviation calculations because of the low number of data points. The measured sensor response and impact force for the previously emphasized tightly grouped sets, corresponding to 15 cm and 20 cm, have signi cantly lower ranges of value than the other measurements. The tight grouping of sensor response and force measurement is likely because the impacts consist of primarily normal components of force, which is the case when the object falls straight down and impacts the sensor at a 90 degree angle. The results of the force characterized collision response method represent a similar result to the previous approximate force method. The relations show sensor response to be a monotonically increasing function of the impact or collision force of the object. Therefore, the sensor response supports the assertions and modeling present in Section 3.6 and achieves collision detection for a wide range of impact forces and applications. Table 5.2: Total result comparison of measured sensor response and impact force object collisions for range of drop heights. The values in parenthesis for the 35 cm sensor response are the values including the outliers. The force measurements did not exhibit outliers. Data Height Force (N) Range (N) sens (GPa) Range (GPa) A 5 cm 49:315 N 14:0956 N 3.0511 0.4282 B 10 cm 64:524 N 21:7028 N 4.4339 0.6959 C 15 cm 84:309 N 6:9066 N 6.6007 0.1204 D 20 cm 113:452 N 1:3143 N 6.9587 0.3212 E 35 cm 220:728 N 35:9005 N 7.9891(6.7089) 0.1606(3.9209) F 50 cm 413:778 N 63:4738 N 6.9828 3.6533 51 Chapter 6 Conclusion In contrast to existing technologies, the detailed sensor design developed in this the- sis shows strong collision sensation for both planar and non-planar surfaces. The pseudo- membrane construction eliminates not only mechanical issues associated with pressure and membrane based sensation but also increases applicability of the PVDF sensing technology. Furthermore, the uniform and consistent response of planar and non-planar applications eliminates hardware specialization needs such that modular collision detection systems can be created. The interface electronics and sensor construction is accomplished with commer- cially producible parts such that retro tting is easily accomplished. The results support the theoretical relation of compressive stress to measured response in the local frame and the sensor measurements provide a monotonically increasing output as a function of the applied collision force. The novel sensor design in this thesis shows strong promise for a robustly applicable collision detection solution for complex robotic arms and non-standard operating environments. 6.1 Results Summary The approximate force testing method utilized for characterizing the sensor show the wide dynamic sensing range of the sensor from levels as low 5 N to values as high as 300 N. The strong correlation and consistent results show the sensing to be capable of detecting collision and limiting harm before signi cant force has occurred to the robotic system or impact object. The force measurement supports and achieved consisten results with the approximate method verifying the measurement. The peak force measurement cover a similar range of 49 N to 419 N. The high level of 419 N clearly demonstrates the sensors collision 52 detection ability at high levels. The deviation of collision for object impacts which do not squarely impact the sensor{still show a collision detection response. 6.2 Future Work Further nite element analysis and strain modeling of the system and substrate is nec- essary to more accurately identify the applicability of the sensor. Speci cally, closed cell foam and other elastic substrates experience stress and strain in a nonlinear manner and more complex modeling and force measurement is required to accurately model the system. Additionally, more extensive testing of substrate options to more accurately characterize the e ects of material properties on sensor response is needed to increase viability of con- struction. The testing was primarily carried out using equivalent size elements and objects to maintain continuity of the experimentation. However, additional testing using a variety of not only di erent sized and shaped elements but also di erent sized and shaped objects would be bene cial. The collisions utilized for testing were limited to impact collision gener- ated from falling objects. Recall from the previously described hazard origins for mechanical danger, di erent types of collisions including pinching scenarios, impact and hold scenarios{a collision in which object is already in contact with the sensor and the robotic system ac- tuates towards the object{ are necessary to ensure the sensor enables the system to meet current international safety standards including ISO 10218 [22, 23] and IEC 60601-1 [28]. Additional algorithm development for advanced or adaptive collision detection utilizing the sampled signal would also bene t the work. In order to satisfy the previously mentioned standards, routines for online calibration and sensor fault testing are necessary. 53 Bibliography [1] A. Mirbagheri, J. Dargahi, F. Aghili, and K. Parsa, \Finite element analysis of a membrane-type piezoelectric tactile sensor with four sensing elements," in 2005 IEEE Sensors, Nov. 2005, pp. 4 pp.{. [2] L. Seminara, L. Pinna, M. Valle, L. Basirico, A. Loi, P. Cosseddu, A. Bon glio, A. Ascia, M. Biso, A. Ansaldo, D. Ricci, and G. Metta, \Piezoelectric polymer transducer arrays for exible tactile sensors," in 2012 IEEE Sensors, Oct. 2012, pp. 1{4. [3] S. Haddadin, A. Albu-Scha er, A. De Luca, and G. Hirzinger, \Collision detection and reaction: A contribution to safe physical human-robot interaction," in IEEE/RSJ International Conference on Intelligent Robots and Systems, IROS 2008, 2008, pp. 3356{ 3363. [4] R. Schulze, T. Gessner, M. Schueller, R. Forke, D. Billep, M. Heinrich, M. Sborikas, and M. Wegener, \Integration of piezoelectric polymer transducers into microsystems for sensing applications," in Applications of Ferroelectrics held jointly with 2012 Eu- ropean Conference on the Applications of Polar Dielectrics and 2012 International Symp Piezoresponse Force Microscopy and Nanoscale Phenomena in Polar Materials (ISAF/ECAPD/PFM), 2012 Intl Symp, 2012, pp. 1{4. [5] A. Yamamoto, S. Nagasawa, H. Yamamoto, and T. Higuchi, \Electrostatic tactile dis- play with thin lm slider and its application to tactile telepresentation systems," IEEE Transactions on Visualization and Computer Graphics, vol. 12, no. 2, pp. 168{177, 2006. [6] I. Fujimoto, Y. Yamada, T. Morizono, Y. Umetani, and T. Maeno, \Development of arti cial nger skin to detect incipient slip for realization of static friction sensation," in Proceedings of IEEE International Conference on Multisensor Fusion and Integration for Intelligent Systems, MFI2003., Jul. 2003, pp. 15{20. [7] E. S. Kolesar and C. S. Dyson, \Object imaging with a piezoelectric robotic tactile sensor," Journal of Microelectromechanical Systems, vol. 4, no. 2, pp. 87{96, Jun. 1995. [8] J. M. Wooten, D. M. Bevly, and J. Y. Hung, \Robust large-area piezoelectric polymer based collision detection sensor," in 39th Annual Conference of the IEEE Industrial Electronics Society (IECON-2013), Vienna, Austria,, Nov. 2013, pp. 1{6. [9] C. S. Brown, R. C. Kell, R. Taylor, and L. A. Thomas, \Piezoelectric materials, a review of progress," IRE Transactions on Component Parts, vol. 9, no. 4, pp. 193{211, 1962. 54 [10] E. Fukada, \Piezoelectricity of wood," Journal of the Physical Society of Japan, vol. 10, no. 2, pp. 149{154, 1955. [Online]. Available: http://jpsj.ipap.jp/link?JPSJ/10/149/ [11] D. Damjanovic, \Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin lms and ceramics," Reports on Progress in Physics, vol. 61, no. 9, p. 1267, 1998. [12] E. Fukada, \History and recent progress in piezoelectric polymers," IEEE Transactions on Ultrasonics, Ferroelectrics and Frequency Control, vol. 47, no. 6, pp. 1277{1290, Nov. 2000. [13] H. Kawai, \The piezoelectricity of poly (vinylidene uoride)," Japanese Journal of Applied Physics, vol. 8, no. 7, pp. 975{976, 1969. [Online]. Available: http://jjap.jsap.jp/link?JJAP/8/975/ [14] J. Cai, X. Fu, and H. Liu, \Design of PVdf stress sensor and its application in the stress test of foam plastic," in 2012 International Conference on Measurement, Information and Control (MIC), vol. 2, 2012, pp. 821{824. [15] Y. Shapiro, A. Wolf, and G. Kosa, \Piezoelectric de ection sensor for a bi-bellows actuator," IEEE/ASME Transactions on Mechatronics, vol. 18, no. 3, pp. 1226{1230, 2013. [16] J. Kymissis, C. Kendall, J. Paradiso, and N. Gershenfeld, \Parasitic power harvesting in shoes," in Second International Symposium on Wearable Computers. Digest of Papers, 1998, pp. 132{139. [17] H. Han, Y. Nakagawa, Y. Takai, K. Kikuchi, and S. Tsuchitani, \PVdf lm micro fabrication for the robotics skin sensor having exibility and high sensitivity," in 2011 Fifth International Conference on Sensing Technology (ICST), 2011, pp. 603{606. [18] T. Sekitani, T. Yokota, and T. Someya, \Ultra- exible, ultra-thin, ultra-sensitive or- ganic pressure sensor system for biomedical applications," in 2012 IEEE Sensors, 2012, pp. 1{3. [19] J. Xia, Z. Xie, H. Fang, T. Lan, J. Huang, and H. Liu, \Collision detection of exible joint manipulator by using joint torque sensors," in Advanced Intelligent Mechatronics, 2009. AIM 2009. IEEE/ASME International Conference on, 2009, pp. 1816{1821. [20] H.-W. Je, J.-Y. Baek, and M.-C. Lee, \A study of the collision detection of robot manipulator without torque sensor," in ICCAS-SICE, 2009, 2009, pp. 4468{4471. [21] L. Seminara, M. Capurro, P. Cirillo, G. Cannata, and M. Valle, \Electromechanical characterization of piezoelectric PVdf polymer lms for tactile sensors in robotics ap- plications," Sensors and Actuators A: Physical, vol. 169, no. 1, pp. 49{58, 2011. [22] Robots and robotic devices - Safety requirements for industrial robots: Part 1: Robots, International Standard Organization Std. ISO 10 218-1, Rev. E, 2011. 55 [23] Robots and robotic devices - Safety requirements for industrial robots: Part 2: Robot systems and integration, International Standard Organization Std. ISO 10 218-2, Rev. E, 2011. [24] Y. Roh, V. Varadan, and V. K. Varadan, \Characterization of all the elastic, dielectric, and piezoelectric constants of uniaxially oriented poled PVdf lms," IEEE Transactions on Ultrasonics, Ferroelectrics and Frequency Control, vol. 49, no. 6, pp. 836{847, 2002. [25] J. Gere, Mechanics of Materials, 5th ed. Brooks/Cole, 2001. [26] F. De Lim, L. B. Tan, C. Quan, and T. Y. Tee, \Whole- eld board strain and displace- ment characterization during drop impact using a single camera dic technique," in 2009 11th Electronics Packaging Technology Conference. EPTC ?09., 2009, pp. 652{655. [27] J. A. Bryson, \Impact response of polyurethane," Ph.D. dissertation, Washington State University, December 2009. [28] Medical electrical equipment - Part 1: General requirements for basic safety and essential performance, International Electrotechnical Commission Std. IEC 60 601-1, Rev. 3.1, December 2012. [29] J. Reddy, An Introduction to Continuum Mechanics. Cambridge University Press, 2008. 56 Appendix A Material Properties A.1 Cauchy Stress Tensor It is important to consider the stress that exists within an object as a state of stress. Furthermore, stress exists as combination of normal and shear stress components which combine to form a tensor. [25] = 2 66 66 4 Te1 Te2 Te3 3 77 77 5 = 2 66 66 4 11 12 13 21 22 23 31 32 33 3 77 77 5 = 2 66 66 4 xx xy xz yx yy yz zx zy zz 3 77 77 5 = 2 66 66 4 x xy xz yx y yz xz yz z 3 77 77 5 (A.1) The Cauchy Stress Tensor is one such state of stress representation often used in contin- uum mechanics. In this case, the stress tensor represents the components of stress for an in nitesimally small point in the material and is de ned as the current force per unit de- formed area, ij. The stress at a point in a three-dimensional continuum can be shown using nine quantities, three per plane, on three mutually perpendicular planes of the point.[29] The stress components for each plane, x,y and z, are combined in Cartesian component form to produce the second-order tensor (A.1).[29] The individual stress vectors, Tei, represent the orthogonal components of stress for each plane. The tensor is utilized to represent complex kinematics and statics in material which exists as a continuous combination of many small in nitesimally small points. The Cauchy Stress tensor is unique because the symmetric properties of the tensor allow for a stress representation of the point using only 6 elements known as a Voigt notation; speci cally, x, y, z represent the normal stresses and the shear stresses are represented by xy, xz, yz, or equivalently xy, xz, or yz. The symmetry of 57 the stress tensor refers to the property: ij = ji where i and j represent x, y, or z. This symmetry is particularly useful in the strain modeling to relate experienced stress of the lm on the surface of the material to stress components of the continuous material. A.2 Engineering Strain Tensor = 2 66 66 4 11 12 13 21 22 23 31 32 33 3 77 77 5 = 2 66 66 4 xx xy xz yx yy yz zx zy zz 3 77 77 5 = 2 66 66 4 x xy2 xz2 yx 2 y yz 2 xz 2 zy 2 z 3 77 77 5 (A.2) Strain components of a material use similar notation to stress and results in a strain tensor. The tensor includes the normal and shear strains. By considering the displacements as small, we can neglect nonlinear terms and the resulting tensor known as the in nitesimal strain tensor is a symmetric second-order tensor in rectangular Cartesian components. The diagonal components of the tensor 11, 22, and 23 represent the normal strains and the o diagonal terms 12, 13, and 23 represent the shear strains. The shear strains are called the engineering shear strains and the tensor becomes the engineering strain tensor in the nal notation [29]. A.3 Geometric Representation of Strain ij = 12( ij o + ji o ) (A.3) An important material property for modeling the sensor, the geometric representation of strain relates the normal strain to the change of length to the previous length of the material. The combination of change in length in i and j direction represented by i and j respectively divided by the original length jo and io yields the normal strain when divided by 2, ij.[25] The geometric property is particularly useful when combined with a symmetric strain tensor 58 because measured surface strains of the lm can be transformed to the normal strain of the material. A.4 Young?s Modulus E = (A.4) The modulus of elasticity introduced by English scientist Thomas Young, Young?s Modulus, relates the tension and compression of a material. Speci cally, the Young?s Modulus of a material is the relation of axial stress to axial strain. The modulus of elasticity is a constant representing the slope of the stress-strain diagram in the linearly elastic region of the graph [25]. The modulus is often used in a form known as Hooke?s law which more clearly relates the stress of the material due to strain, shown in equation (A.5). = E (A.5) The typical units for E are pascals because strain, , is a unit less measurement and the SI unit for stress, , is pascals or gigapascals. A.5 Shear Modulus of Elasticity G = (A.6) The shear modulus of elasticity is a similar relation to previously discussed Young?s Modulus. The modulus represents the slope of the linear region of the shear stress-strain curve and is frequently found in the Hooke?s law in shear, shown in Equation (A.7). = G (A.7) 59 The shear modulus of elasticity is typically expressed in pascals such that the shear stress, , is in the SI unit for stress when the modulus, G, is multiplied times the unit less shear strain, [25]. 60