List of Figures

Figure 1.1 Core prognostic frameworks in a PHM system. Source: based on IEEE (2017).
Figure 1.2 Framework diagram for a PHM system. Source: based on CAVE3 (2015).
Figure 1.3 Graph of the exponential CDF with λ = 3.
Figure 1.4 Graph of the Weibull CDF with β = 1.2 and η = 5.
Figure 1.5 Graph of the exponential PDF with λ = 3.
Figure 1.6 Graphs of gamma PDFs.
Figure 1.7 Graphs of Weibull PDFs.
Figure 1.8 Failure rates of Weibull variables.
Figure 1.9 Failure rates of gamma variables.
Figure 1.10 Failure rate of the standard normal variable.
Figure 1.11 Failure rate of the lognormal variable with μ = 0 and σ = 1.
Figure 1.12 Logistic failure rate with μ = 0 and σ = 1.
Figure 1.13 Gumbel failure rate.
Figure 1.14 Log‐logistic failure rate with μ = 0.
Figure 1.15 Integration domain.
Figure 1.16 High‐level block diagram of a PHM system.
Figure 1.17 A framework for CBM for PHM. Source: after CAVE3 (2015).
Figure 1.18 Taxonomy of prognostic approaches.
Figure 1.19 Example of an FFP signature – a curvilinear (convex), noisy characteristic curve.
Figure 1.20 Ideal DPS transfer curve superimposed on an FFP signature.
Figure 1.21 Ideal DPS, degradation threshold, and functional failure.
Figure 1.22 Normalized and transformed FFP and DPS transformed into FFS.
Figure 1.23 Ideal FFS – transfer curve for CBD.
Figure 1.24 Variability in DPS transfer curves.
Figure 1.25 FFS transforms of the DPS plots shown in Figure 1.23.
Figure 1.26 FFS and prognostic information.
Figure 1.27 FFS transfer curve exhibiting distortion, noise, and change in degradation rate.
Figure 1.28 Example plots of an ideal RUL and ideal PH.
Figure 1.29 Example plots of an ideal SoH transfer curve and PH accuracy.
Figure 1.30 Example of RUL with an initial‐estimate error of 100 days.
Figure 1.31 Random‐walk with Kalman‐like filtering solution for a high‐value initial‐estimate error.
Figure 1.32 Random‐walk with Kalman‐like filtering solution for a low‐value initial‐estimate error.
Figure 1.33 Example of FFS data exhibiting an offset error, distortion, and noise.
Figure 1.34 Utility function of price.
Figure 1.35 Utility function of expected lifetime.
Figure 1.36 Shape of the objective function.
Figure 2.1 Block diagram showing three approaches to PHM. Source: based on Pecht (2008).
Figure 2.2 Precision and complexity: relative comparison of classical PHM approaches.
Figure 2.3 Model‐based approach to development and use.
Figure 2.4 Model‐use diagram. Source: based on Medjaher and Zerhouni (2013).
Figure 2.5 A framework for CBM for PHM (CAVE3 2015).
Figure 2.6 Transition diagram.
Figure 2.7 Example of a fault tree showing an RTD fault and an RTD‐usage model.
Figure 2.8 Example plots of Weibull distributions.
Figure 2.9 A special system structure.
Figure 2.10 HALT result – 30 of 32 FPGA devices failed (Hofmeister et al. 2006).
Figure 2.11 Family of failure curves, failure distribution, and TTF.
Figure 2.12 Diagram of data‐driven approaches.
Figure 2.13 A special neural network.
Figure 2.14 Comparison of model‐based (PoF) and data‐driven prognostic approaches.
Figure 2.15 Relative comparison of PHM approaches – PoF, data‐driven, and hybrid.
Figure 2.16 Example diagram of a heuristic‐based CBM system using CBD‐based modeling.
Figure 2.17 Diagram comparison of model‐based and CBD‐signature approaches to PHM.
Figure 2.18 Simplified diagram of a switch‐mode power supply (SMPS) with an output filter.
Figure 2.19 Unreliability plots for three models for capacitor failures (Alan et al. 2011).
Figure 2.20 Example of the output of a SMPS.
Figure 2.21 Example of a ringing response from an electrical circuit to an abrupt stimulus.
Figure 2.22 Relationship of prognostic specifications (PD and PDα) to RUL and PH.
Figure 2.23 Simulated change in resonant frequency as filter capacitance degrades.
Figure 2.24 Experimental change in resonant frequency as filter capacitance degrades.
Figure 3.1 Diagram of classical and CBD prognostic approaches for PHM systems. Source: based on Pecht (2008).
Figure 3.2 Functional block diagram for CBD signature data and processing flow.
Figure 3.3 Example of CBD containing feature data and noise (FD + Noise).
Figure 3.4 Example of signals at an output node of an SMPS.
Figure 3.5 Example of a damped‐ringing response (Judkins and Hofmeister 2007).
Figure 3.6 Modeling a damped‐ringing response. Source: based on Judkins and Hofmeister (2007).
Figure 3.7 Example of a CBD signature: FD is the resonant frequency of a damped‐ringing response.
Figure 3.8 CBD signature and levels for SMPS lot 1 (top) and SMPS lot 2 (bottom).
Figure 3.9 Functional block diagram for FFP signature and processing flow.
Figure 3.10 FFP signatures using a fixed value and a calibrated value for nominal frequency.
Figure 3.11 FFP signature, calibrated value for nominal frequency, failure threshold at 0.6 and 0.7.
Figure 3.12 FFS signatures for FL = 0.6 (top) and FL = 0.7 (bottom).
Figure 3.13 DPS and FFP signatures for data shown in Figure 3.11.
Figure 3.14 FFP and DPS Showing FL = 0.4 and FL = 0.5.
Figure 3.15 DPS‐based FFS (FL = 0.65) and FFP‐based FFS (FL = 0.7).
Figure 3.16 Examples of signatures: decreasing (top) and increasing (bottom) slope angles.
Figure 3.17 Other examples of signatures: decreasing (top) and increasing (bottom) slope angles.
Figure 3.18 Simulated CBD‐based signature: FD = CBD − NM.
Figure 3.19 Differences: experimental and simulated signatures.
Figure 3.20 Simulated (top) and comparison of experiment (bottom) FFP signatures.
Figure 3.21 Simulated DPS (top) and experimental DPS (bottom).
Figure 3.22 Simulated FFS from simulated (top) and from experimental DPS (bottom): FL = 0.65.
Figure 3.23 Illustration of point‐by‐point FNL comparison.
Figure 3.24 Illustration of total FNLE comparison.
Figure 3.25 Example plot of FFP‐based and DPS‐based FNLi.
Figure 3.26 SMPS output and extracted damped‐ringing response.
Figure 3.27 CBD signature and FFP signature.
Figure 3.28 FFP‐based FFS and DPS.
Figure 3.29 DPS‐based FFS.
Figure 3.30 Procedural diagram for producing a DPS‐based FFS.
Figure 3.31 Linear DPS (left side) and nonlinear DPS (right side) data plots.
Figure 3.32 Simulated CBD plots using nonlinear and linear degradation rates.
Figure 3.33 Comparison of ideal DPS to DPS from data for a nonconstant degradation rate.
Figure 3.34 Node‐based framework for supporting failure progression signatures.
Figure 4.1 Block diagram for offline modeling of CBD signatures.
Figure 4.2 Flow diagram for developing signature models.
Figure 4.3 Power function #1: increasing curves, decreasing slope angles.
Figure 4.4 Power function #1: increasing curves, increasing slope angles.
Figure 4.5 Power function #2: decreasing curves, decreasing slope angles.
Figure 4.6 Power function #2: decreasing curves, increasing slope angles.
Figure 4.7 Power function #3: increasing curves, vertically asymptotic, dPi < P0.
Figure 4.8 Power function #4: decreasing curves, vertically asymptotic, dPi < P0.
Figure 4.9 Power function #5: increasing curves, horizontally asymptotic.
Figure 4.10 Power function #6: decreasing curves, horizontally asymptotic.
Figure 4.11 Power function #7: increasing curves, slightly curvilinear, decreasing slope angles.
Figure 4.12 Power function #7: increasing curves, slightly curvilinear, increasing slope angles.
Figure 4.13 Power function #8: decreasing curves, slightly curvilinear, decreasing slope angles.
Figure 4.14 Power function #8: decreasing curves, slightly curvilinear, increasing slope angles.
Figure 4.15 Power function #9: increasing curves, vertically asymptotic, dPi < P0.
Figure 4.16 Power function #9: increasing curves, horizontally asymptotic, dPi < P0.
Figure 4.17 Power function #10: decreasing curves, vertically asymptotic, dPi < P0.
Figure 4.18 Power function #10: decreasing curves, horizontally asymptotic, dPi < P0.
Figure 4.19 Exponential function #11: increasing curves, vertically asymptotic.
Figure 4.20 Exponential function #12: decreasing curves, vertically asymptotic.
Figure 4.21 Exponential function #13: increasing curves, horizontally asymptotic.
Figure 4.22 Exponential function #14: decreasing curves, horizontally asymptotic.
Figure 4.23 Simulated FFP signatures: power function #1.
Figure 4.24 Simulated DPS from FFP signatures: power function #1.
Figure 4.25 Simulated FFP signatures: power function #3.
Figure 4.26 Simulated DPS from FFP signatures: power function #3.
Figure 4.27 Simulated FFP signatures: power function #5.
Figure 4.28 Simulated DPS from FFP signatures: power function #5.
Figure 4.29 Simulated FFP signatures: power function #7.
Figure 4.30 Simulated DPS from FFP signatures: power function #7.
Figure 4.31 Simulated FFP signatures: power function #9.
Figure 4.32 Simulated DPS from FFP signatures: power function #9.
Figure 4.33 Simulated FFP signatures: exponential function #11.
Figure 4.34 Simulated DPS from FFP signatures: exponential function #11.
Figure 4.35 Simulated FFP signatures: exponential function #13.
Figure 4.36 Simulated DPS from FFP signatures: exponential function #13.
Figure 4.37 Simulated FFP and DPS signatures: power function #1 for n = 2.0.
Figure 4.38 Simulated FFP and DPS signatures: power function #3 for n = 2.0.
Figure 4.39 Simulated FFP and DPS signatures: power function #5 for n = 1.5.
Figure 4.40 Simulated FFP and DPS signatures: power function #7 for n = 0.75.
Figure 4.41 Simulated FFP and DPS signatures: power function #9 for n = 0.25.
Figure 4.42 Simulated FFP and DPS signatures: exponential function #11 for P0 = 100`.
Figure 4.43 Simulated FFP and DPS signatures: exponential function #13 for P0 = 150.
Figure 4.44 Example plots: FFS (top) and FFP and DPS (bottom): power function #1.
Figure 4.45 Example plots: FFS (top) and FFP and DPS (bottom): power function #5.
Figure 5.1 Example of a non‐ideal FFP signature and an ideal representation of that signature.
Figure 5.2 Plots of a family of FFP signatures and DPS transfer curves.
Figure 5.3 Plots of a curvilinear FFP, the transform to a linear DPS (top), and the transform to an FFS (bottom).
Figure 5.4 Offline phase to develop a prognostic‐enabling solution of a PHM system.
Figure 5.5 Diagram of an online phase to exploit a prognostic‐enabling solution.
Figure 5.6 Plot of a non‐ideal CBD signature data: noisy ripple voltage, output of a switched‐mode regulator.
Figure 5.7 Example plots: non‐ideal FFP signature data (top) and transformed DPS data (bottom).
Figure 5.8 Example plots: non‐ideal and ideal FFS data (top) and FNL (bottom).
Figure 5.9 Example plots: non‐ideal and ideal FFS data (top) and FNL (bottom) after using a NM of 3.0 mV.
Figure 5.10 Example plot: non‐ideal ADC transfer curve.
Figure 5.11 Example plots: ideal input and non‐ideal output from an ADC.
Figure 5.12 Temperature (a), voltage (b), and current (c) plots.
Figure 5.13 Temperature‐dependent (a) and temperature‐independent (b) plots of calculated resistance.
Figure 5.14 Example of a damped‐ringing response.
Figure 5.15 Sampling diagram for Example 5.8.
Figure 5.16 ADC example: saw‐tooth input, sampling, digital output value.
Figure 5.17 Simulated data before (top) and after (bottom) filtering of white (random) noise.
Figure 5.18 Degradation signature exhibiting a change in shape.
Figure 5.19 Experimental data: temperature measurements for a jet engine.
Figure 5.20 Differential signature from temperature measurements for each of two engines.
Figure 5.21 Temperature data and differential signatures: four engines on an aircraft.
Figure 5.22 Composite differential‐distance signature.
Figure 5.23 Example of a noisy FFP signature.
Figure 5.24 Example of a noisy FFP signature: calculated nominal FD value, reduced NM value.
Figure 5.25 Example of a smoothed (3‐point moving average) FFP signature.
Figure 5.26 Example of a DPS from a smoothed FFP signature.
Figure 5.27 Example of an FFS from a smoothed FFP signature.
Figure 5.28 Example of a {FNLi} plot from a smoothed FFS.
Figure 5.29 Example random‐walk paths and FFS input.
Figure 5.30 Example plots of input FFS data and adjusted FFS data.
Figure 5.31 Example of a {FNLi} plot from an adjusted FFS.
Figure 5.32 Ripple voltage: plots of an unsmoothed (top) and smoothed (bottom) FFP signature.
Figure 5.33 Ripple voltage: plots of an unsmoothed (top) and smoothed (bottom) FFS data.
Figure 5.34 Smoothed FFS: for NM = 3.0 mV (top) and for NM = 2.0 mV (bottom).
Figure 6.1 Core prognostic frameworks in a PHM system. Source: after IEEE (2017).
Figure 6.2 A framework for CBM‐based PHM. Source: after CAVE3 (2015).
Figure 6.3 Random walk with Kalman‐like filtering solution for a high‐value initial‐estimate error.
Figure 6.4 Random walk with Kalman‐like filtering solution for a low‐value initial‐estimate error.
Figure 6.5 Block diagram showing three approaches to PHM.
Figure 6.6 Model development and use.
Figure 6.7 Diagram: model‐based and CBD‐signature approaches to PHM.
Figure 6.8 Example diagram: heuristic‐based CBM system using CBD‐based modeling.
Figure 6.9 Procedural diagram for producing a DPS‐based FFS.
Figure 6.10 Examples of CBD, FFP, DPS, and DPS‐based FFS.
Figure 6.11 DPS‐based FFS and FFP‐based FFS.
Figure 6.12 FNL plots for the FFS shown in Figure 6.11.
Figure 6.13 Plots of a family of FFP signatures and DPS transfer curves.
Figure 6.14 Offline phase to develop a prognostic‐enabling solution for a PHM system.
Figure 6.15 Online phase to exploit a prognostic‐enabling solution.
Figure 6.16 Multiple temperature signals: before and after differential‐distance conditioning.
Figure 6.17 Extracted FD from fusing differential‐distance conditioned CBD.
Figure 6.18 Block diagram of an example EMA subsystem.
Figure 6.19 Offline modeling and development diagram.
Figure 6.20 Design and analysis diagram.
Figure 6.21 Continual (top) and periodic (bottom) sampling.
Figure 6.22 Period‐burst sampling: sampling period (TS) and burst period (TB).
Figure 6.23 Example of power supplies and EMA subsystems.
Figure 6.24 SMPS output showing ripple period (TR) and sampling period (TS).
Figure 6.25 Damped‐ringing response caused by an abrupt load change.
Figure 6.26 Example of burst of burst sampling.
Figure 6.27 Example of alerts issued using an unrealistic PHM system clock and/or polling method.
Figure 6.28 Steps in test data due to low‐resolution fault injection.
Figure 6.29 Diagram of a test bed to inject faults into a power supply.
Figure 6.30 Diagram of a test bed to inject faults into an EMA.
Figure 6.31 Sampled‐ and windowed‐phase currents: no load (top) and extra load (bottom).
Figure 6.32 FFP signature of fault‐injected SMPS.
Figure 6.33 Sampled‐phase currents: no load (top) and extra load (bottom).
Figure 6.34 Current magnitude: no degradation (high), degraded transistor (reduced amplitude).
Figure 6.35 Shifted levels due to a degraded power‐switching transistor.
Figure 6.36 Illustration of using peak positive and negative threshold values.
Figure 6.37 Special rms: threshold and truncation.
Figure 6.38 Special rms applied to three phase currents.
Figure 6.39 FFP signatures due to loading.
Figure 6.40 Smoothed FFP signatures due to loading.
Figure 6.41 FFS: EMA loading.
Figure 6.42 Smoothed FFP signatures due to winding faults.
Figure 6.43 FFS: EMA winding.
Figure 6.44 Phase A currents: transistor fault in the positive Phase A portion.
Figure 6.45 Phase B currents: transistor fault in the positive Phase A portion.
Figure 6.46 Phase C currents: transistor fault in the positive Phase A portion.
Figure 6.47 Peak‐RMS currents: transistor fault in the positive Phase A portion.
Figure 6.48 Close‐up of the EMA motor winding.
Figure 6.49 FFP: H bridge fault, sum of both halves of the Phase A current.
Figure 6.50 Smoothed FFP: H bridge fault.
Figure 6.51 Smoothed FFS: H bridge fault.
Figure 6.52 Block diagram for an example of a robust PHM system.
Figure 6.53 Architectural block diagram for a node definition.
Figure 6.54 Example of a system node definition.
Figure 6.55 Block diagram of an example node definition.
Figure 6.56 NDEF: node status.
Figure 6.57 NDEF: sampling specifications.
Figure 6.58 NDEF: alert specifications.
Figure 6.59 NDEF: Special Files specifications.
Figure 6.60 NDEF: feature‐vector framework – (a) primary, (b) smoothing, (c) FFP‐DPS transform.
Figure 6.61 NDEF: Prediction Framework.
Figure 6.62 NDEF: Performance Services/Graphics.
Figure 6.63 NDEF: Input & Output Files.
Figure 6.64 NDEF: Checkpoint Library & File Name.
Figure 6.65 NDEF: Device Driver Program ID and Units of Measure.
Figure 6.66 NDEF: Other Program IDs.
Figure 6.67 NDEF: End of Definition.
Figure 6.68 NDEF updates to support node second SMPS (node 60).
Figure 6.69 NDEF updates to support EMA 1 (node 51).
Figure 6.70 Illustration and relationship of PH to BD, sample time, RUL, and EOL.
Figure 6.71 Relationship of degradation times and an FFS.
Figure 6.72 Illustration of the uncertainty of determining exactly when functional failure occurs.
Figure 6.73 Uncertainty: prognostic distance.
Figure 6.74 CBD at 1‐hour sampling (top) and at 24‐hour sampling (bottom).
Figure 6.75 Family of failure curves, failure distribution, and TTF.
Figure 6.76 Diagram of an initial estimate for PD.
Figure 6.77 Comparison: FFS and plots of RUL and PH estimates.
Figure 6.78 Prognostic bus (log file).
Figure 6.79 SoH plot from DXARULE.
Figure 6.80 Output file for SMPS using ARULEAV.
Figure 6.81 Plots of the RUL, PH, and SoH estimates produced by ARULEAV.
Figure 6.82 Output file for SMPS using DPS‐based FFS and ARULEAV.
Figure 6.83 Plots of the RUL, PH, and SoH estimates using DPS‐based FFS and ARULEAV.
Figure 6.84 CBD, FFP, and FFS for EMA node 51 (friction/load).
Figure 6.85 EMA (load) plots: RUL, PH, and SoH estimates using DPS‐based FFS and ARULEAV.
Figure 6.86 CBD, FFP, and FFS for EMA node 61 (winding).
Figure 6.87 EMA (winding) plots: RUL, PH, and SoH estimates; DPS‐based FFS and ARULEAV.
Figure 6.88 CBD, FFP, and FFS for EMA node 62 (power transistor).
Figure 6.89 EMA (transistor) plots: RUL, PH, and SoH estimates; DPS‐based FFS and ARULEAV.
Figure 6.90 PHM: high‐level control and data flow.
Figure 6.91 Initialization: system nodes.
Figure 6.92 System alerts, part 1.
Figure 6.93 System alerts, part 2.
Figure 6.94 Example of a GUI for a PHM system. Source: Ridgetop (2018).
Figure 7.1 Example of a broad view of an ecosystem.
Figure 7.2 Example of a five‐level model for health solutions.
Figure 7.3 Example of alerts for SoH at or below 25%.
Figure 7.4 Example of alerts for a damage‐detection approach.
Figure 7.5 MTTF, TTF, and PITTFF0: CBD signature and failure distribution.
Figure 7.6 Same MTTF for different failure distributions and signatures.
Figure 7.7 Failure plots with average values for TTF and PD = PITTFF0.
Figure 7.8 Estimated PD and actual PD: power supply (top) and EMA load (bottom).
Figure 7.9 Estimated PD and actual PD: EMA winding (top) and EMA transistor (bottom).
Figure 7.10 RUL and PH plot for PITTFF0 = 4800 and for PITTFF0 = 2290.
Figure 7.11 RUL and PH plots for SMPS: before and after PITTFADJ = 2.0.
Figure 7.12 RUL and PH plots for EMA 51: before and after PITTFADJ = 2.0.
Figure 7.13 RUL and PH plots for EMA 61: before and after PITTFADJ = 2.0.
Figure 7.14 RUL and PH plots for EMA 62: before and after PITTFADJ = 2.0.
Figure 7.15 Plots: test results for six power supplies (top) and 12 EMAs (bottom).
Figure 7.16 Bathtub curve showing three regions, MTBF, and a prognostic trigger point.
Figure 7.17 Possible relationship of bathtub curve to failure distribution and MTTF.
Figure 7.18 Multiple instances of CBD signatures and trigger points.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset