You are on page 1of 8

Improved Rotor Track and Balance Performance

Using an Expert System


Eric Bechhoefer
NRG Systems (formerly Goodrich SIS), Hinesburg, VT

Austin Fang
Sikorsky Aircraft Corporation, Stratford, CT 06615

Daniel Van Ness


Goodrich SIS, Vergennes, VT 05491
AbstractA major source of vibration in the helicopter is derived from the main rotor dynamics and imbalance. Excessive vibration can cause premature wear to the aircraft and rotating components. Additionally, vibration causes pilot fatigue that in itself can be a safety issue. While great improvements have been made in the rotor track and balance (RTB) techniques, it still requires a skilled maintainer. Systems such as the Goodrich IVHMU can give good solutions. However, it requires a high degree of user expertise in selecting which RTB control adjustments should be active for the current helicopter measured vibration. Selection of RTB control adjustments or solution strategies is based on the type of imbalance/vibrations that are being generated by the main rotor. A system which recognizes the current vibration state can be trained to use the best controls and solution strategy. Such an expert system would improve the quality of the RTB solution and in general reduce the overall workload of the maintainer. We present a framework for a Bayesian classifier expert system. The classifier is based on a multiple hypothesis testing and demonstrates the ability to deliver an improved RTB adjustment over a generalized solution strategy. Rotor Track and Balance; Expert System; Bayesian Classifier

lateral vibration may also be the result of an out of track condition. For example, if one blade is flying high and producing more lift, it will also lag its normal position due to higher drag force. This induces an effective mass imbalance due to the aerodynamic forces imposed on the blade being out of position [1]. This also moves the center of mass of that blade inboard compared to others. Vertical vibrations are caused by blade track variations. A blade lifting the helicopter in one quadrant of rotation and losing lift in another quadrant causes this phenomenology. In forward flight, the flapping motion of the blade (due to increased lift of the advancing blade) can additionally generate a 2 per revolution vibration [2]. II. ROTOR TRACK AND BALANCE ADJUSTMENTS

A typical rotor system has: hub weights, pitch control rods (PCR) and tab adjustments. Some manufactures may have two (inboard and outboard) tabs and/or tip weights. The following discussion will apply to a 4-blade rotor system, but is applicable to any number of blades. For a 1 per revolution adjustment, only two blades (or hub) positions need to be affected. For example, adding weight to one hub position is the same as removing weight from the opposite blade. PCR and tab adjustments will change the track of a blade, which will also change the blades lead/lag (and hence, lateral vibration). A PCR changes the pitch at the root of the blade while a tab affects the aerodynamic loading of a blade outboard on the blade. PCR effectiveness is constant with aircraft speed, while a tabs effectiveness increases with speed. It is generally agreed that a track reduction only (flat track) solution does not necessarily give the best vibration [2]. The end result of a RTB maintenance event should be the smoothest possible ride and not solely based on track.

I.

INTRODUCTION

There are many sources of vibration from rotating components on the helicopter, but we will focus on the main rotor blade. There are two main directionality of vibration associated with the main rotor: lateral and vertical. Lateral vibration is the result of unequal distribution of mass in the rotating disk [1]. This imbalance can be the result of manufacturing error, erosion, absorption of water into the blade, or poor assembly techniques. A The information and methods contained herein are protected by active and pending Patents of Simmonds Precision Products, Inc., a wholly-owned subsidiary of the Goodrich Corporation.

Real blade adjustments are quantized/rounded off estimates of a Fourier adjustment. The vibration vector itself is a complex number, where the phase of the adjustment is referenced to a known blade/position on the aircraft. For example, on the UH-60, each blade is referenced by color: black (0 degrees), yellow (90 degrees), blue (180 degrees) and red (270 degrees) relative to the forward centerline of the helicopter. Consider a simple example where the cockpit vertical (derived vertical measurement with phase considerations) 1 per rev vibration at 90 knots is 57o at 1.2 inch per second (ips). The 2 per rev vibration are 38o at 0.16 ips. We will use a PCR adjustment. The vibration coefficients of the PCR adjustment for 1 per rev (for example purposes) is 132o at 0.13 ips/notch and for 2 per rev is 338o at 0.054 ips/notch. We wish to reduce the vibration to zero. We can model this as such that the current vibration is V, the vibration coefficients is X, and the PCR adjustment is A. Note that these represent discrete Fourier (DFT) values. The 1 per rev vibration is orthogonal to the 2 per rev vibration. This requires solving 2 sets of equations to derive the DFT values for the 1 per rev adjustment and a DFT value for the 2 per rev adjustment. The solution is simply: A = V*X-1 For 1 per rev, we have: 2.4 8.9i = (0.65 + 1.0i)*(-0.09 - 0.1i)-1 and for 2 per rev: 1.6 +2.6i = (0.13 + 0.1i)*(0.05 + 0.03i)-1 The real blade adjustment associated with the DFT adjustment can be found by multiplying by the inverse discrete Fourier transform. For a four blade rotor, the transform is trivial: the four equivalent real adjustments are: For 1 per rev: Pick One Tupple black +2.4, yellow blue -2.4, yellow black +2.4, red blue -2.4, red -8.9 -8.9 +8.9 +8.9 (1)

Here it is important to note that there is no reason to implement 4 blade changes. We can simply subtract one blade value from all of the other blades to implement 3

Figure 1 One and Two Per Rev Adjustment in Time adjustments (equivalent to changing the DC offset, or blade conning angle). Note that we implement only whole PCR notch adjustments. The four equivalent total real blade adjustments are then:
Black 0 8 5 -10 Yellow -8 0 -3 -18 Blue -5 3 0 -14 Red 10 18 14 0

Of course, there are maintainers, or other preferences, which would prompt one adjustment over the other. Some selection criteria might be: Minimize total PCR change. Adjustment 1 has a total change of -3 PCR vs. 29, 16, -42 clicks. A large accumulation in PCR in any give direction could over time change the rigging of the helicopter. This in turn could limit the maximum rotor RPM in an autorotation: effectively allowing a slow creep out of limits. Minimum number of blades adjusted. There may be a minimum allowable change, such as having no tab adjustments less than 4 mil, or no PCR adjustment less than 2 clicks. In this case, Adjustment 2 would have only 1 adjusted, -11 on the Yellow blade. Minimizing track split. Each of the PCR adjustment will change track split. Given the current track split, we could estimate the track after the adjustment and select the adjustment set that both minimizes vibration and track.

For 2 per rev: black -1.6, yellow 1.6, blue -1.6, red 1.6 See figure 1 and a visual representation of the adjustment The total real blade adjustment is the superposition of both the 1 and 2 per rev solution.
Black 0.9 Yellow -7.3 Blue -4.0 Red 10.5

In general, a RTB solution is more complex than the prior example. Rotor smoothing is performed over at least two sensors (cockpit lateral and cockpit vertical, but could also include cabin lateral, cabin vertical, cabin longitudinal, etc) and multiple flight regimes (e.g. ground, hover, 80 kts, 125 kts, 145 kts, etc). Additionally, there may be different objectives for the optimization: minimize vibration, minimize track, and minimize both vibration and track. III. RTB ALGORITHMS

a tab solution would result in a poor adjustment (e.g. the tab adjustment would be large due to the low sensitivity coefficient when trying to control a lateral vibration with track). While a more advanced solution strategy may be attempted, it would be a discrete, non-linear problem which would be inherently difficult to solve [8]. Alternatively, an Expert System decision module was developed with Sikorsky Aircraft Corporation to evaluate the current helicopter vibration states. The expert system selects which adjustments would be most appropriate for the current vibration vector used in the RTB event. The expert system partitions the helicopter track and vibration universe into an n dimensional solution space. Each dimension has an appropriate set of adjustments and solution strategy assigned to it. The selection of the best adjustment can be modeled as a Bayesian decision problem, where the expert system algorithm picks the most similar decision space in reference to the observation space. This can be viewed as multiple hypothesis testing, where the decision algorithm develops evidence from the observation space, and if the evidence is large, rejects a null hypothesis (a default decision space) in favor of an alternative hypothesis (a more attractive decision space). A. Hypothesis Testing: Bayesian Classifier Consider a binary system, where the set of observations corresponds to a default decision space (Weight, PCR, Inboard Tab, Outboard Tab, Track, and Minimum Vibration (3 blade adjustment)) or some alternative decision space (Weight, Inboard Tab, and Minimum Adjustment (2 blade adjustment)). After sampling the observation space (vibration and track data for sensors and regimes), two possible decision outcomes can be assigned: Ho or Ha (note that without loss of generality, there could be m > 2 decision spaces, in which case this would be a multiple hypothesis test). By convention, Ho is called the null hypothesis and in this example, corresponds to the default decision space). The observation is a parameterized measurement of a system, which is a function of some probabilistic law. This reduces the hypothesis-testing problem to one of deciding which hypothesis represents truth, based on the measurement z. The range of z is the observation space. The decision problem consists of partitioning this observation space into two regions, zo (default) and za (alternative). When z falls within zo one says Ho is true and when z falls within za, ones say that Ha is true. When a decision is made that is incorrect (say Ha when in fact Ho is true), an error has occurred. An error for RTB would represent selection of a sub-optimal decision space for RTB and implementation of inappropriate adjustments. The end goal of the hypothesis-testing is to create a decision region that minimizes the error with the given observation space z. In the general case, the observation

The objective of a RTB maintenance event is to apply adjustment(s) such as weights, PCRs or tabs to give the smoothest possible ride. We can define the adjustment and initial helicopter vibration in the Fourier domain, as demonstrated. Since the adjustment will typically contain more than one adjustment type, we will group define a vector, A, representing the vector of adjustments, and similarly, a vector V, representing the sensor vibrations (e.g. cockpit lateral, cockpit vertical, cabin lateral, etc.) and helicopter regimes (ground, hover, 80 kts, etc.). We will define the vibration after the adjustment at U, which is the same dimension as V. The generalized problem is then: U = f(V,A). The exact optimization is typically proprietary to the RTB equipment manufacturer. VEMP (Vibration Management Enhancement Program) optimizes the reduction in vibration via a neural network [5]. While the neural network is capable of modeling non-linear relationships, the application for RTB was found to be linear [4]. The Goodrich IVHMU system models vibration linearly [6, 7], such that: U = V XA, where X is the vibration coefficient matrix. This models the relationship between an adjustment and its effect on vibration at a given sensor and regime. Depending on the objective function, this could be solved using least square (e.g. finding A such that the sum(U2) is a minimum. The Goodrich system objective function minimizes AEAT subject to the constraint that | V-XA| < , where E is a vector of weights proportional to the maximum adjustment size and is the total system error [6, 7]. The Goodrich system additionally has two solution strategies. For low vibration and smaller track split, only shaft order one (SO1) is solved for. For higher vibration (SO1 and SO2) or large track splits, a solution using both SO1 and SO2 is solved for, resulting in PCR and TAB adjustments on three blades (for a four blade helicopter). Note that these solution strategies fundamentally do not address the issue of which adjustments are to be used for a given RTB event. For example, if track split is low, and vibration is primarily lateral in the ground regime, a weights only control adjustment solution is appropriate. Using a default solution strategy will give a good balance (facilitated by large degree of freedom), but all of the adjustments will be small (which a maintainer will not appreciate). Alternatively, applying

space consists of a set of parametric observations z = (z1, z2, ,zn) with some joint probability density function p(z1, z2, ,zn). B. The Bayesian Classifier for Normal Distributions Under many circumstances, the Normal distribution is a valid model of the data. In the case of an n dimensional observation space, due to the Central Limit Theorem, the Normal distribution should be the default model. For the generalized, n dimension decision space,

hypothesis, and i = 2 to n alternative hypothesis are tested. If h(y) < 0, the default (null) hypothesis is taken. Otherwise, the index with the greatest value of h(y) is the most likely decision space. A. Configuration: Determination of m and P In the assumed Gaussian model, each mi represents the mean value of the vibration vector associated with an adjustment set. Each Pi is the weighting applied to the error term (e.g. normalized distance) between the measured vibration and the adjustment set i. This constitutes the a priori configuration representing the knowledge of an expert RTB maintainer. This was done by using Sikorsky supplied flight data and known successful decision spaces and subsequent solutions on the S-92A. Initially one needs to define the adjustment space for the test aircraft model, in this case the S-92A. The S92A is a medium helicopter with a 4-Blade rotor. Each rotor blade has 4 possible adjustments: Weights, PCR, Inboard Tabs, and Outboard Tabs. The Goodrich RTB algorithm additionally has two solutions strategies: minimum vibration (optimization of 1 and 2 per rev) or minimum adjustment (optimization only 1 per rev). This gives the following decision space (Table 1).
TABLE I. Index 1 ALL POSSIBLE SOLUTION STRATEGIES FOR THE S-92A RTB Wgt X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X PCR X X X InTab OutTab MinVib X X X X X X X X X X X X X

H :Z=V 0 p( 0)= 1 z|H


and

(2)

n 2

1/2

e p[ zT1z] x 0
(7)

H :Z=V+m 1 pz|H)= 1 ( 1
(8) as:

(2)

n 2

1/2

e p z-m 1( ) x ( ) 0 z-m
T

The normalized distance squared between z and m


2 d ( m z m = ) 1 ) z ( T

(9)

The log likelihood ratio test then is:


2 2 1 d 0 d1 2

ln 0 1 H0
> <

H1

1/ 2

2 3 4

(10)

5 6 7 8 9 10 11 12 13 14 15 16 17 18

where is the covariance and || is the determinant of the covariance. This states that if the normalized distance between z and m0 is greater than the normalized distance between z and m1, then the null hypothesis is rejected [9]. IV. EXPERT SYSTEM FOR RTB1 The previous derivation shows the optimality is akin to minimizing the distance between a given decision space and the measurement vector. In order to simplifying the implementation, it was assumed that the covariance is diagonal. This allows one to define inverse variance (normally called information) as preference, Pi. The decision algorithm simplifies to:

h )=(ym 2P (ym)2P + (y ) 1 1 2 2 P 2 l n <0 > P 1 2


1

(11)

19 20 21

For the case where there are multiple decision spaces (i > 2), the first decision space is taken at the null
1

Patent Pending, Simmonds Precision Products, Inc.

22 23 24 25 26 27

X X X

X X X X X X X X

Log ratio

180

-44

-462

-288

33

Note that a weight only solution will always use two blades, thus is defined as a Minimum Adjustment solution with Minimum Vibration deselected. From this initial decision space, we partitioned the problem into 6 solutions spaces (Table II):
TABLE II. Index 1 2 3 4 5 6 X X Wgt X SUBSET OF ALL POSSIBLE SOLUTIONS: S-92A RTB PCR X X X X X X X X X X X OutTab InTab X MinVib X X

Where H0 represents index 1 from Table II, H1 represents index 2 from Table II, etc. For H0 (Wgt, PCR, Inboard Tab using the minimum vibration solution strategy) to be selected, the decision rule requires the log likely ratio to be negative for all tests. Since this is not the case, the expert system chooses the index for the test with the greatest log likely ratio: H1, or solution space 2: a Minimum Vibration (3 blade) solution with PCR and Outboard Tabs. Note that the next most attractive solution is the full adjustment (Wgt, PCR, Inboard and Outboard Tabs, with Minimum Vibration). See table III, column 6). The recommended adjustment is:
TABLE IV. Blade Weight PCR OBTab IBTab Blue 0 0 0 0 SCENARIO 1, RECOMMENDED ADJUSTMENT Red 0 -10 8 0 Black 0 6 -13 0 Yellow 0 -9 0 0

The maximum vibration was reduced to less than 0.4 ips (Overhead lateral moving from 0.5 to 0.2 ips in all regimes, with the mean 1per rev vibration over six sensors, all regimes, going from 0.28 ips to 0.19 ips) and the track split was reduced to 30 mm. See figure 2. The default solution strategy resulted in the following adjustments:
TABLE V. Blade Weight PCR OBTab IBTab Blue 0 13 -8 -27 SCENARIO 1, DEFAULT ADJUSTMENTS Red -31 0 17 -11 Black -32 15 -13 -23 Yellow 0 0 0 0

This subset of possible solutions will cover most conditions likely to be encountered by a maintainer. There are three Minimum Adjustment based solutions that would be appropriate in low vibration and/or small track split situations (e.g. rotor tuning), and three Minimum Vibration solutions to correct larger 1 or 2 per rev vibration or large track splits. A supervised learning technique was implemented using a dataset of 58 RTB operations from different 18 aircraft. Each operation was allocated into the appropriate adjustment sets. Each operation contained regime (ground, hover, 120, 140 and 150 kts), vibration harmonic (1 through 8 per rev, but only 1 and 2 per rev were used), vibration sensor channels (cockpit vertical, pilot longitude, roll, pilot lateral, cabin vertical and overhead lateral) and track data. The mean value and inverse covariance was then calculated for each adjustment set based on the population. B. Scenario 1: Large Track Split in Flight The pilot reports a large track split (max of 80mm) which is speed dependent (e.g. track split increases with speed). The measured vibration is relatively low, the maximum being less than 0.5 ips. Using equation (11), the log likely ratios are (Table III):
TABLE III. Test Ho v. H1 LOG LIKELY RATIOS OF DECISION SPACES Ho H2 v. Ho H3 v. Ho H4 v. Ho v. H5

The total vibration reduction was not significantly lower (mean 1per rev vibration over six sensor and all regimes, going from 0.28 ips to 0.17 ips) while the track split was almost twice as great (59 mm) as the expert system solution of 30 mm. Finally, the default adjustment set required 10 vs. 5 blade touch operations to implements. Note that the reduction of track split was a function of the optimization reducing 2 per rev. There are two other optimization strategies that could be employed to reduce flat track: a combination of reduction of vibration (1 per rev) and track (instead vibration for the 2 per rev optimization) or a flat track solution. As an example, the flat track solution resulted in effectively no change in 1per rev vibration (mean 1per rev vibration over six sensor and all regimes, going from 0.28 ips to 0.27 ips, with only a marginal

improvement in 2per rev vibration, 0.27 to 0.24 ips), with a track split of 13 mm.

solution as a trade off comparison to the Expert System developed solution. The Wgts/PCR Minimum Adjustment solution was:
Blade Weight PCR Blue 153 0 Red -73 -13 Black 0 -27 Yellow 0 0

This adjustment is similar to the Minimum Vibration solution. The difference being that the track split increased to 106 mm (which is large), but with a similar reduction in the mean vibration of 0.14 ips. The Weight alone solution gave:
Blade Weight Blue -31 Red -59 Black 0 Yellow 0

Figure 2 Comparison of Track Split from ES and Default Solution From a maintainer perspective, the expert system solution is superior in that its fewer adjustments reduces vibration to almost the lowest achievable level with the default adjustment, while reducing the track split to an acceptable level. C. Scenario 2: Large Ground Vibration After a repair of the blade deicing system, the pilot reports a large vibration during the ground turn. The HUMS system indicates the pilot lateral vibration is 1.0 ips, and cockpit roll vibration is 0.27 ips. The other sensor report lower vibration levels (0.2 ips) and the 2 per rev are seen as 0.1 ips or less. The acceptance test procedures for the S-92A state: The primary measurements used for ground balance are cockpit roll (A-B) vibration and pilot floor lateral vibration. The Main Rotor Ground Balance limit for roll is 0.20 ips; floor lateral limit is 0.25 ips. Because ground vibration dynamics are influenced by the loading of 2 dampers, namely the tires (pressure) and strut dampers, the optimization weighting is focused on the acceptance criteria, cockpit roll and pilot lateral. The expert system recommends a Weights and PCR solution using the Minimum Vibration solver.
Blade Weight PCR Blue 155 27 Red -74 3 Black 0 0 Yellow 0 16

The pilot lateral vibration decreased to 0.34 ips while the cockpit roll decreased to 0.22 ips. The mean 1per rev vibration over all sensors actually increased from 0.49 ips to 0.51 ips. While the track split remained small, the post adjustment pilot lateral and roll vibration exceeded the acceptance criteria (see figures 3 and 4).

Figure 3 Polar Plots for Vibration for Different Solution Strategies It must be noted that this is a large adjustment. In practice, vibration of this level will require inspection and a functional check flight. Additionally, it would be expected that this may require two or more adjustment cycles. In this example, once the ground vibration is reduced (as verified by a rotor turn), the maintainer would again runt the RTB algorithm. In this case, to pull in/reduce the track split. Once that was accomplished, the helicopter would be moved into forward flight regimes to determine 1 and 2 per rev vibration. If these were within the acceptance criteria, the aircraft would be signed off. Previously it had been stated that a flat track did not necessarily imply a low vibration. That being said, there are limits for acceptable track split: the expert system solution is preferable to the Wgts/PCR Minimum Adjustment solution in that the track split is 20 mm less.

The pilot lateral vibration was reduced to 0.06 ips, and the cockpit roll vibration decreased to 0.08 ips, which is below the acceptance criteria (see figure 3). The track split went from 30 mm to 54 mm (see figure 4). The mean vibration went from 0.49 ips to 0.14 ips. This can be compared with a Weights/PCR Minimum Adjustment solution and a Weights only

optimal adjustments by sorting through the decision space to determine the most appropriate adjustment set. The Bayesian classifier was shown to be able to selecting a better solution strategy and adjustment set than the default adjustment set. The expert system demonstrated its ability to incorporate OEM or advanced user knowledge through a priori configuration data. This improvement to the RTB algorithm will increase the overall ability of maintainers to perform rotor tuning and reduce the total maintenance cost associated with operating a helicopter. REFERENCES
[1] Robinson, M., Rotor Track and Balance Theory, Aircraft Maintenance Technology, February 1999. [2] Johnson, L., History: Helicopter Rotor Smoothing:, http://www.dssmicro.com/theory/dsrothst.htm. [3] FM 1-514: Fundamentals of Rotor and Power Train Maintenance Techniques and Procedures, Headquarters, Department of the Army, Washington DC., 5 April 1991. [4] Miller, N., A Comparison of Main Rotor Smoothing Adjustments Using Linear and Neural Network Algorithms, Air Force Institute of Technology, Wright-Patterson Air Force Base, OH., 2006 [5] Giurgiutiu, V., Grant, L., Grabill, P., Helicopter Health Monitoring and Failure Prevention Through Vibration Management Enhancement Program 54th Meeting of the Society for Machinery Failure Prevention Technology, May 14, Virginia Beach, VA [6] Bechhoefer, E., Power, D., IMD HUMS Rotor Track and Balance Techniques IEEE Aerospace Conf., Big Sky, 2003. [7] Bechhoefer, E., Ventres, C., Reducing Vibraiton Using QR Decomposition and Unconstrained Optimization US Patent 6,567, 757 B2, May 20, 2003 [8] Avriel, M., Golany, B., Mathematical Programming for Industrial Engineers, Marcel Dekker Inc., New York, 1996, pg 385-409.M. Young, The Technical Writer's Handbook. Mill Valley, CA: University Science, 1989. [9] Fukunaga, K., Introduction to Statistical Pattern Recognition, Academic Press, San Diego, 1990. [10] Hagan, M., Demuth, H., and Beale, M., Neural Network Design, Tomson Learning, Bangkok, 1996

Figure 4 Track Split for Different Solution Strategies V. CONCLUSIONS

We have shown that the proposed Bayesian classifier expert system develops a better adjustment than using default settings. This has the advantage of improving RTB maintenance practices. With this result, two additional questions need to be addressed. First, given the processing speed available, why not solve using all 27 possible solutions (Table 1) and pick the best solution. Second, one might ask why would the user not implement a neural network (NN) instead of Bayesian classifier? With regard to processing speed, in addition to incurring the cost of processing all 27 solutions, the decision support technology required to select the best solution is identical to the problem just solved. Namely, it would require configuration to define what a good solution is and an algorithm to search the solution space and quantify the best adjustment. In response to algorithm selection, it has been our experience that it is difficult to certify a neural network (NN) within the context of certification standard DO178B because, in general, the solution to finding the weights in the perceptron is not deterministic. It is noted that the Bayesian classifier has similar structure to a single level, feed forward perceptron [10]. This suggests that the performance (e.g. correct classification) of the NN would not be any better than a Bayesian Classifier. In fact, it is likely that the NN would require more training sets than may be available. Sikorsky spurred the development of the RTB expert system in identifying the need to improve the overall user experience of RTB,. The RTB algorithm is a powerful tool, with a number of options useful in reducing the vibration on a helicopter. However, those options require a high level of training and on-hands experience to take advantage of that capability. The expert system reduces the chance of implementing non-

Eric R. Bechhoefer is the lead system engineer at NRG Systems. A former naval aviator, Dr. Bechhoefer recently joined NRG from the aerospace industry. Dr. Bechhoefer has 16 patents and over 50 papers related to condition monitoring of rotating equipment. Austin Fang is a senior engineer at Sikorsky Aircraft Corporation. After graduating Cornell University with an M.S. in Mechanical Engineering in 2005, Austin joined the Noise, Vibration, and Harshness group as a dynamics engineer. Austin holds a fixed wing private pilot certificate and enjoys working on condition monitoring applicable to rotorcraft. Daniel Van Ness is a systems engineer at Goodrich Corporation. Daniel graduated from the University of Notre Dame with a Ph.D. in Aerospace Engineering, and has since joined the HUMS engineering group at Goodrich Sensors and Integrated Systems.

You might also like