Next Article in Journal
Experiments of Lift-Bending Response on a Slender UAV Wing Structure with Control Surface under Extreme Flow Turbulence
Next Article in Special Issue
Combination of Advanced Actuator Line/Disk Model and High-Order Unstructured Finite Volume Solver for Helicopter Rotors
Previous Article in Journal
Design of Flyby Trajectories with Powered Gravity and Aerogravity Assist Maneuvers
Previous Article in Special Issue
Air and Structural Loads Analysis of a 5-Ton Class Rotorcraft in a Pull-Up Maneuver Using CFD/CSD Coupled Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibration Qualification Campaign on Main Landing Gear System for High-Speed Compound Helicopter

by
Antonio Chiariello
1,*,
Carmine Carandente Tartaglia
1,
Maurizio Arena
2,*,
Vincenzo Quaranta
1,
Giovanni Bruno
1,
Marika Belardo
1 and
Martina Castaldo
2
1
Italian Aerospace Research Centre (CIRA), Via Maiorise snc, 81043 Capua, CE, Italy
2
Loads & Performance Unit, Magnaghi Aeronautica of MA Group Company, Via Galileo Ferraris 76, 80146 Napoli, NA, Italy
*
Authors to whom correspondence should be addressed.
Aerospace 2024, 11(2), 130; https://doi.org/10.3390/aerospace11020130
Submission received: 18 December 2023 / Revised: 24 January 2024 / Accepted: 31 January 2024 / Published: 1 February 2024
(This article belongs to the Special Issue Advances in Rotorcraft Dynamics)

Abstract

:
Vibrations in helicopters have strong implications for their performance and safety, leading to the increased fatigue of components and reduced operational efficiency. As helicopters are designed to land on several types of surfaces, the landing gear system dissipates the impact on the ground and maintains stability during landing and take-off. These vibrations can arise from a variety of sources, such as aerodynamic loads, mechanical imbalances, and engine instabilities. In the present work, the authors describe the vibration qualification process of the main landing gear tailored to fast helicopters within the Clean Sky 2 Racer program. The method entails devising preliminary load sets that deform the structure in its key excited mode shapes to assess stresses and address the experimental campaign. A full-scale prototype model is then tested for sine sweep and random vibrations as per the Airbus Helicopter requirements in order to reach the final qualification and acceptance stage. Although the discussion centers on a landing gear structure, the described process could be extended to other critical equipment as well.

1. Introduction

1.1. Vibration Characterization in Aerospace Field

The prediction and optimization of vibration are essential steps to determine the contributing sources of disturbance on the propeller vehicle’s performance. Propeller aircraft are generally subjected to the dynamic sources provided by the engine and airborne fluctuations. Consequently, dynamic analysis has received attention since the first phases of the aircraft design process [1,2,3,4]. The ground vibration test (GVT) is generally carried out by assessing the aircraft structural strength to verify the compliance with dynamic and aeroelastic requirements [5]. The technical–scientific literature on the vibration aspects of aero-structures is very extensive. However, most of the few existing works are purely based on numerical models [6], with no experimental validation or evaluation. The employment of composite materials has led to the development of more complex aircraft designs and consequently has imposed further challenges for GVT. This increase in structural complexity has resulted in a new scenario, where several research efforts [7,8,9,10,11] have been undertaken in order to reduce the testing time and cost, as well as improving its accuracy. The study in [12] focused on the details of the GVT of an all carbon-fiber-reinforced polymer (CFRP) composite unmanned aerial vehicle (UAV). Modal parameters including natural frequencies, mode shapes, and damping coefficients were numerically estimated for the full aircraft in a free-free condition and then compared to the experimental vibration characteristics measured by means of a shaker-table approach. Previous studies have performed finite element (FE) simulations and static and vibration testing of the wing assembly [13,14] and fuselage [15]. Gupta et al. [16] conducted a GVT on a flying wing aircraft within the NASA project entitled ‘Performance Adaptive Aeroelastic Wing’ (PAAW); in such a context, two tested configurations were designed (i.e., Sköll and Hati). Some studies have reported GVTs involving non-linear modal analysis and the inherent non-linear concerns [17,18,19]. The landing systems have long been the object of considerable interest regarding the study of the dynamic stability of the aircraft on the ground, especially for shimmy- and brake-induced vibrations [20]. Shimmy vibration is an unsteady behavior of the landing gear system during either takeoff or landing, endangering the safety and ride of the aircraft. Energy input for this type of vibration is provided by the kinetic energy of the forward motion of the aircraft [21]. As a result, this energy initiates a self-excited type of vibration in the wheels that may lead to an uncontrolled condition for ground maneuvering. An attempt was made in [22] to develop a dynamical model for the aircraft nose gear, to investigate its transient response to the lateral deviations and shimmy angles. Variations in different design parameters—namely, the energy absorption coefficient of the shimmy damper and the location of the gravity center of the landing gear—were studied. Both increasing the torsional damping and optimizing the stiffness distribution of the NLG can effectively improve the anti-swing performance, preventing the shimmy [23,24,25]. Several instability suppression strategies have been investigated in recent decades by developing multi-disciplinary formulations and the innovative design of active vibration absorbers [26,27,28,29,30,31,32,33]. Friction-induced vibrations are a major concern in the aircraft braking mechanism. Two significant non-linear phenomena have been recognized in the disk brakes during the taxi and landing phases: squeal and whirl. The other primary modal interactions were chatter and gear walk. Brake design parameters such as the working pressure, disk stiffness and geometry, and friction coefficient could be optimized to improve the brake roll conditions [34,35,36,37,38]. While modal interactions on the ground have been extensively discussed, a discussion of the in-flight conditions is lacking. The present authors aim to discuss the vibratory dynamics of the main landing gear (MLG) when exposed to the operative excitations of the aircraft. Despite the use of standard design and testing tools, the article presents novel activity within the certification of landing gear systems for the latest generation of high-speed vehicles.

1.2. RACER Main Landing Gear

The RACER MLG is based on a direct cantilever architecture, with an integrated nitrogen-oil shock absorber (S/A) consisting of one stage: a separator piston divides the oil and nitrogen phases. The MLG is sustained laterally by a side-brace actuator (SBA), which can extend and retract the leg and also lock it both in the down and up positions. In such a way, the actuator combines the dual function of the hydraulic power unit and structural bracing. The MLG system, including the hydraulic actuator and wheel and brake (W&B) assembly, is detailed in Figure 1.

1.3. Scope of the Research

The landing gear structure is designed to withstand the static and cyclic loads encountered during all the landing cases and the ground. Additionally, the dynamic loads comprise low-frequency sine waves and background random noise. The first are generally the excitation harmonics of the propeller source, namely the blade passing frequencies (BPFs), while the random noise—of a lower amplitude—is instead a consequence of the aero-structural interaction and turbulent boundary layer (TBL). A vibration analysis is usually essential when the expected acceleration levels may fall next to the structural natural frequencies. This paper describes a finite element (FE)-based procedure for the identification of the vibration levels of the MLG with regard to the in-flight inertial loads. The present study arises from the need to explore the system’s structural behavior due to the narrow distance between the helicopter (H/C) tonal loads and the predominant modal deformations. In the framework of the same research project, the authors have already performed the characterization of the vibration levels of the MLG composite door as a function of the propeller operating tonal loads [38]. Hence, to perform a structural strength analysis of the landing gear and conduct qualification tests, a numerical evaluation is performed to identify the main dynamic parameters (resonance frequencies, generalized masses). The central themes focus on the modeling strategies, dynamic analysis at harmonic and random loads, and final laboratory testing activities. Due to the various methods available to simulate the two mentioned load entities, an equivalent static load is considered as an engineering combination of them. The optimization of computational times for FE models based on a large number of nodes and elements is in fact a crucial topic currently addressed in the engineering literature. The approximation to a quasi-static type of analysis has the significant advantage of reducing the computational costs compared to an advanced analysis in the time or frequency domain. Transient analyses require the numerical integration of the differential equations governing the structural dynamics for all DOFs of the system. Spectral methods based on the extrapolation of eigenvectors possess the great advantage of replacing the dynamic equilibrium with a set of algebraic equations [39,40]. The normal modes and the corresponding modal mass distribution are employed to evaluate a possible equivalent static load and investigate the relevant stress effects in the resonance condition. In other words, the generalized mass calculation plays a key role in identifying the dynamic loads acting on the landing gear. The qualified modal base of the MLG system could represent proof of its compliance with airworthiness and safety regulations.

2. Materials and Methods

2.1. Numerical Model Design

The 3D FE models are conceived to be fully representative of the detailed MLG components; see Figure 2. The material characteristics are defined in Table 1. Structural FE analysis was carried out using 3D FEM modeling (in the Altair Hypermesh environment). Mesh modeling included the adoption of solid (ctet), rigid (rbe), and contact (cgap) elements, according to the formal classification of the MSC Nastran® handbook [41] (MSC Software 2021, Newport Beach, CA, USA). Contact analysis allows us to simulate the load interaction among the sub-components. The main characteristics of the 3D FE model are summarized in Table 2.

2.2. FEM Boundary Conditions

The external constraints are representative of the actual attachments of the MLG leg and actuator vibration test fixture; see Figure 3. The wheel and brake component is assumed as a lumped mass (conm2, [41]) including the equivalent inertial properties; see Table 3.
Some details of the contact modeling strategy are indicated in Figure 4.

2.3. Sine-on-Random Loading Conditions

The vibration performance and endurance test profiles from Airbus Helicopters combine a sequence of sine waves corresponding to the H/C tones and random noise; see Table 4. The numerical calculations were therefore performed with respect to the harmonics of the main rotor (MR) and lateral rotor (LR) and the random profile (10–2000 Hz); see Figure 5.

2.4. Computational Methodology for Dynamic Loads

The load conditions that are applied to the MLG structure and retraction actuator in the longitudinal (x), lateral (y), and vertical (z) directions are computed by carrying out the following equations. The total dynamic load combines both low-frequency harmonic force and a broadband random spectrum relying on the main participating normal mode shapes; see Figure 6.
  • The harmonic sinusoidal load can be introduced as a factor of gravitational acceleration (ain):
F s = m g a i n Q
where Q = 1/2ξ is the transmissibility at resonance occurrence and mg is the generalized mass actually participating in the modal deformation.
2.
The random load to be applied is approximated by means of the Miles equation [43]:
x ¨ r m s = π 2 f n i Q W u i ¨ ( f n )
F r = m g x ¨ p e a k
where fn is the resonance frequency and W U ¨ ( f n ) is the relevant auto-power spectral density (ASD) of the random enforced acceleration. The 3σ values of the random load factors are then applied considering an equivalent static peak load ( x ¨ p e a k = 3 x ¨ r m s ).

3. Numerical Results

3.1. MLG Normal Mode Analysis

The first stage was to characterize the dynamic properties of the MLG leg and re-traction actuator (natural frequencies, modal shapes, and participating mass) on the basis of FE models. The key mode shapes were identified in the bandwidth [0, 50 Hz]—considered significant for the dynamic displacement entity—by a linear modal solver [41]. For this reason, only the modes expected to be the most representative of a cantilevered telescopic LG (flexural deformations in fore and aft and lateral direction) and of a slender actuator strut (typical bending) are provided; see Figure 7. The following Table 5 indicates the corresponding natural frequencies and modal masses calculated in the range of interest. Both the first MLG leg and hydraulic actuator mode shape fall close to the range of the second sine tone (i.e., 1/rev LR), which will be subsequently investigated for the structural analysis aspects. The other elastic deformations not reported are beyond the excitation bands and therefore not contemplated for the dynamic load evaluation.

3.2. Model Validation against Experiment

Considering the complexity of the main landing gear (MLG) and the severity of the qualification loads, with the aim of verifying the reliability of the numerical model, the MLG was subjected to trial tests before carrying out the whole vibration test campaign. For this reason, the MLG was subjected to sine sweep at 0.5 g from 10 Hz to 2000 Hz at 1 octave/min in all three directions. With reference to the directions defined in Section 4 (for excitation in the X and Z directions), the MLG was mounted on the shaker slip table; see Figure 8a,b. The response of the system was recorded at the same points as in the tests carried out along the Z axis (presented in detail in the Section 4).
In order to perform the trial test in the Y direction, the MLG was mounted on the shaker head expander; see Figure 9.
For a landing gear system, the wheel certainly represents one of the most stressed elements. Furthermore, the numerical results highlighted that the first resonance frequency for the MLG wheel was close in frequency to the second sinusoidal tone imposed by the tailoring of the standard and representative of the blade passage frequencies.
The results in Table 6 refer to the accelerometer on the wheel system named AC4 and positioned as outlined in Section 4. In Figure 10, it is possible to see how this point was monitored during the experimental tests, as done in the numerical analysis (Figure 3). Although a modal analysis is the most appropriate process to identify all the dynamic characteristics, a sine-on-random vibration test is adequate to perform a survey of the natural frequencies and relevant amplification factors Q. The results demonstrate that the first natural frequencies measured are very close to those estimated numerically but above all confirm their closeness to the second excitation tone. A well-addressed numerical investigation was therefore considered of key relevance. Furthermore, the actual amplification factors Q are determined by the conservative preliminary choice.
The positive outcome of the trial test made it possible to verify that the numerical model was able to predict the behavior of the system in a conservative manner. The difference between the numerical and testing resonance frequencies may be due to the stiffness modeling approach of constraints, which will be tuned in a subsequent detailed phase of this research activity. Although the test was only required for the MLG system without the wheel, given the positive results of the trial test, along the axis considered most critical, i.e., the Z axis, the test was carried out with the wheel system. The other two axes, considered less critical, were tested without a wheel system, to avoid over-testing. The same analysis was also performed for the extended SBA; see Figure 11. For this EUT, the reference system is defined in Section 4, and since the most stressed point was the one positioned halfway along the extension of the actuator (see Figure 12), the results obtained during the trial test for this point are shown in Table 7. Moreover, in this case, the results demonstrated the ability of the numerical model to test the system in a conservative manner.
The trial phase developed as described so far allows us to highlight potential critical issues in the system (and/or in the test setup) and to adopt the necessary corrective measures aimed at successfully carrying out the qualification tests. Vice versa, as in this case, the positive results of the trial test phase initiated the vibration test campaign with reasonable confidence in the success of the campaign itself.

3.3. FE Calculation of Quasi-Steady Structural Loads

The modal parameters allow for the evaluation of the stress distribution when the structure is subjected to the acceleration sets defined in Table 4. A structural damping coefficient ξ equal to 0.025 is assumed as an engineering choice for metallic assemblies with joints as per ref. [44]; a dynamic amplification factor Q = 1/2ξ = 20 is therefore used for the sine load estimation at the resonance condition. The equivalent static load is applied on the most excited point as established by the normal mode analysis. In this sense, the point at the wheel location would represent intuitively the MLG location with the maximum deformation due to the cantilevered architecture and the W&B lumped inertial effect, as demonstrated numerically in Figure 7a,b. For the actuator, a uniform inertial load is assumed to be applied, which induces a bending load according to the first flexural mode shape; see Figure 7b. The two assemblies—the MLG leg and actuator—are analyzed separately according to the qualification test procedure, which will be discussed in the next paragraph. The MLG equivalent loads with components (Fx, Fy, Fz) are applied at the point with the maximum modal displacement (αx, αy, αz) and calculated for the n mode shapes according to Equation (4):
F x F y F z i = i = 1 n ( F s + F r ) α x α y α z i
The normalized displacements for the three modes of interest are the following:
  • αx = 0.98; αy = 1.00; αz = 0.68; for the first MLG mode shape;
  • αx = 1.00; αy = 0.68; αz = 0.48; for the second MLG mode shape;
  • αx = 1.00; αy = 0.00; αz = 0.00; for the actuator first mode shape.
The equivalent deformations assuming the above (x,y,z) coefficients are given in Figure 13, providing a comparable pattern to the normal modes of Figure 7.
The Von Mises (VM) stress contours (combining principal stresses and shear) represented on the deformation patterns are represented in Figure 14; the details of the most loaded item are addressed in Figure 15. The results include both the MLG leg mode shapes in order to obtain a first conservative evaluation of the stress state. Tensile stresses are more likely to be a source of fatigue failure; therefore, the Max Principal (Max P) is worthy of verification in order to aid the identification of potential critical hotspots. Vibrations can lead to increased wear, particularly in areas where contact pressure could be significant; the oscillating loads with consequent bending deflections could in fact stress considerably the hydraulic actuator, whose nominal design is essentially a tie-rod/strut concept. For this reason, the Min Principal (Min P) stress needs to be investigated differently for the telescopic LG legs already normally sized against the high drag load conditions (spin-up/spring-back).
The FE analyses allowed us to better classify the critical areas to be monitored during the test and, in particular, for the post-test surface checks. The numerical results reported represent an overall view of all load conditions that precisely superimpose the static loads in the three axes simultaneously. Table 8 recapitulates the maximum stress components (endurance acceleration levels) with the preeminent load direction. The Von Mises stress levels for the most loaded parts are below the yield strength (tension fty and compression fcy) of the materials [42]. The most critical load for the MLG leg is in the x direction; in the same way as a cantilever beam, the bending moment trend of the LG structure is a dual function of the shear forces at the free end and the leg extension (arm). The load path passes through the torque links, which—given the eccentricity with respect to the wheel center—transfer a torque to the main fitting. The main fitting is in fact highly stressed at its interface lug with the torque links. Due to the rotation induced by the drag force (x dir.) around the S/A axis, both the torque links tend to deflect laterally with tension stress (Max P) near their common hinge. Moreover, the telescopic S/A could be perceived to be sustained at the two internal bearings (lower and upper). The stress magnitude on the sliding rod—occurring at the lower bearing—depends mainly on the S/A stroke (fully extended condition corresponds to the most conservative case). The loaded parts of the retraction actuator are located close to the mid-station: the scheme can be simplified to a double-pinned beam based on a normal load. Consequently, particular attention was given to the piston rod bending (Max P) and its contact interaction with the end fitting. The bearing stress (Min P) of the end fitting occurs, in fact, at the seal cavity and is emphasized by the global actuator deflection.

3.4. Vibration Preliminary Fatigue Analysis

The fatigue analysis is performed starting with the stress analysis results. The Goodman criterion is, in particular, used to obtain the damage for each fatigue cycle. Fatigue life assessment is performed with the available S-N discrete data, assuming a stress ratio R = −1.0 (pure sinusoidal oscillating load), and their corresponding best-fit curves, based on the data and equations reported in [42]. The materials’ S-N curves are processed with the following reduction factors:
-
surface finish factors [45];
-
surface treatment factor (confidential);
-
reliability factor [46].
The cycles required ni are calculated according to the first MLG resonance frequency and the actuator, as well as with respect to the dwell time to be applied during the test; see Table 9.
The total damage in each load sequence is the sum of Di, according to the Miner rule, as indicated in the Table 10. The damages are well below 1.0, except for that of the MLG sliding rod, which has a higher state of stress.

4. Vibration Qualification Process

4.1. Equipment under Test

The equipment under test (EUT) is reported in Table 11 and consists of the main landing gear (MLG), as shown in Figure 16, and the related side brace actuator (SBA), as shown in Figure 14. To simulate the operating conditions, the actuator was tested in both configurations: retracted (Figure 17a) and extended (Figure 17b). The EUT is designed to be installed on the Airbus Helicopter Racer compound helicopter high-speed demonstrator [47] currently being developed as part of the Clean Sky 2 research program [48]. The tests were conducted in the CIRA’s Space Qualification Laboratory [49].

4.2. Test Specification Tailoring

The purpose of the vibration test campaign was to prove the resistance of the test articles to qualification loads. In order to identify an efficient testing strategy, it was necessary to understand which loads were the most representative, and this led to an analysis of the aircraft configuration in relation to the flight profile.
The aircraft is able to perform vertical takeoff and landing like a helicopter, thanks to the main rotor and its capability to achieve the performance of the fixed-wing aircraft with propellers for the cruise phase. Hybrid aircraft like the racer are not regulated by any standards, and this necessitated a dedicated tailoring phase of RTCA-DO-160-F [50] in order to identify the most representative configuration of the real system and to test the equipment in a conservative condition in terms of the mechanical loads and test setup. The category identified was U2—Unknown Helicopter Frequencies. Each EUT was then tested with sine-on-random vibration loads. Test levels were calculated according to the aircraft source excitation frequency values outlined in Figure 18 [51], by following RTCA-DO-160-F (depending on the H/C vibration zone). The results contained in these tables are directly related to some information classified as confidential, and, for this reason, they are not shown in this paper.
Based on the previous analyses, each set of EUTs was assigned to the following zones:
  • NLG and DBA are placed under the cockpit and therefore belong to area A;
  • MLG and SBA are placed under the wing and therefore belong to zone B.
The duration of each test, as well as the random vibration levels, was the same for each set of EUTs.
A sine sweep test is not required within RTCA-DO-160-F. However, in order to detect any dynamic response variation or structural fault, highlighted by frequency peak changes or amplitude response modifications that could arise from the applied vibrational test loads, it was chosen, in line with what is established by the ECSS standard [52], to carry out a sine sweep at 0.5 [g] in the frequency range of interest, i.e., 10–2000 [Hz]. The sweep rate was chosen, in line with the sine-on-random tests, at 1 octave/min.
The test vibration sequence included the following steps:
  • sine sweep at 0.5 g from 10 Hz to 2000 Hz at 1 octave/min;
  • sine-on-random performance test level for a minimum of 10 min;
  • sine-on-random endurance test level for a minimum of 20 min;
  • repeat performance test level for a minimum of 10 min;
  • repeat sine sweep at 0.5 g from 10 Hz to 2000 Hz at 1 octave/min.
The sine-on-random test is a combined test where several sine sweep tones are superimposed on the random background signal. The result of the RTCA-DO-160-F tailoring phase is summarized in Table 12, where the vibrational loads to which it was subjected are indicated for each set of EUTs. The data are classified as confidential, and, for this reason, the reported sweeping sine tone amplitude values have been normalized to the maximum value. Meanwhile, the random values appear to be as per the standard [52] and therefore are reported.
According to RTCA-DO 160-F, the sinusoidal frequencies during the sine-on-random tests vary at a logarithmic sweep rate not exceeding 1 octave/min from 0.8∙fn to 1.2∙fn, where fn represents the sinusoidal frequencies of the test spectrum [50].

4.3. Test Setup

The tailoring phase made it possible to identify the qualification loads of the landing gear system. These loads are introduced into the system by the shaker through the fixtures. The design of the fixtures must ensure that, in the frequency range of interest, i.e., [20–2000 Hz], the fixtures do not dynamically amplify the input provided by the shaker, and the constraint conditions must be as realistic as possible. For this reason, the attachment points on the fixture were equipped, as in real operating conditions, with movable bearings and bushings. The orientation of the EUT with respect to the shaker plane (normal to the direction of the gravitational force) is irrelevant, since the effects of gravity are negligible on the dynamic response of the system. When we make this assumption, we consider that the structural configuration is evaluated relative to the static equilibrium point and there is essentially no significant deformation of the structure due to the effects of gravity. This was demonstrated on the FE model, in which a negligible value (less than 1 mm in both vertical and horizontal positions) was measured with the EUT unloaded, i.e., when the only force acting on the EUT itself was the force of gravity; see Figure 19.
Since the choice of the relative EUT/shaker orientation was equivalent, it was preferred to position the EUT horizontally (as shown in Figure 20) in order to minimize the unbalancing couples linked to the application of a non-barycentric dynamic load, which would have been balanced by the shaker bearings.
In order to verify that, in the test range, the fixture does not behave by amplifying the response provided by the shaker [51,52], for each set of EUTs, the fixtures were subjected to sine at 0.5 g in the test frequency range (10–2000 Hz, RTCA-DO-160-F Environmental Conditions and Test Procedures for Airborne Equipment, 6 December 2007) by monitoring the responses of the EUT attachment points to the fixtures themselves.
Due to the mass and dimensions shown in Table 12, each subsystem was tested individually on the vibration system coupled with the slip table (X and Z axes excitation) or with the head expander (Y axis excitation) by means of a mechanical fixture, as shown in Figure 16 and Figure 17.
Since the purpose of this paper is to provide a comparison between the experimental and the numerical results obtained when focusing on the MLG along the Z axis (see Figure 20a) and on the SBA in an extended configuration along the X direction (see Figure 20b), because they consider the elements in the most critical configuration, below, we refer to these sets of EUTs.

4.3.1. Test Setup: MLG

The control of the tests performed on the MLG was carried out on the average [50] of three accelerometers (CTR1, CTR2, and CTR3) located in the attachment points of the EUT with the fixture, as shown in Figure 21. Furthermore, four measuring accelerometers (AC1, AC2, AC3, and AC4) were positioned on the EUT, as shown in Figure 22.
For excitation in the Z direction, the EUT was mounted on the shaker slip table, as shown in Figure 23.

4.3.2. Test Setup: SBA in Extended Configuration

The control of the tests performed on the SBA in the extended configuration was carried out on the average of two accelerometers (CTR1 and CTR2) located in the attachment points of the EUT with the fixture; see Figure 24. Furthermore, four measuring accelerometers (AC1, AC2, AC3, and AC4) were positioned on the EUT, as shown in Figure 25.
Figure 26 represents the test setup of the SBA extended for the X direction.

5. Test Results

According to the test sequence reported previously, for each test article, a sine sweep before and after the sine-on-random sequences was performed for each axis. The sine sweeps were performed in order to detect any dynamic response variation or structural fault, highlighted by changes in the frequency of the peaks or amplitude response modifications, which could arise from the applied test loads [53].
In order to compare the results obtained numerically with the experimental ones, the resonance searches performed before (the blue curve) and after (the red curve) the sine-on-random sequences along the Z axis of the aircraft reference system, which corresponded to the Z axis of the MLG reference system and the X axis of the SBA reference system, are shown in Figure 27, Figure 28, Figure 29 and Figure 30 for the MLG and in Figure 31, Figure 32, Figure 33 and Figure 34 for the SBA in the extended configuration. Moreover, for the purposes of this paper, it is necessary to describe the responses of the accelerometers placed at the points in the structure considered most critical, which, for this reason, were subjected to in-depth numerical analyses before proceeding with the tests. These points, in particular, are
  • AC4 for the MLG;
  • AC2 for the SBA.
For these points, the FRFs are shown in Figure 30 and Figure 32, respectively.
Since the extended configuration is the most critical, results will be reported only for this configuration, although the same tests were also performed in the retracted configuration.
The overlapping of the two curves evidenced no significant peak variations, both in terms of amplitude and frequency, as highlighted in Table 13, where the percentage shift in the response of the AC4 accelerometer before and after the S/N test along the Z axis direction is reported for the MLG. Similar results were also obtained for the other accelerometers in all three directions and for each EUT. For resonance searches, the success criterion established was, for modes with an effective mass greater than 10% [52], as indicated in Table 5,
  • less than 5% in frequency shift;
  • less than 20% in amplitude shift.
Table 13 highlights that the resonance search success criterion was satisfied.
Furthermore, at the end of the test sequence, the MLG and the SBA, in each configuration, were visually inspected and no evidence of oil leaks or structural failure was found [50], satisfying the test success criterion as established by RTCA-DO-160-F. The excitation in the X and Y directions showed similar results. Moreover, successive functional tests carried out by Magnaghi Aeronautica proved the correct functioning of the landing gear system and its capability to extend and retract without any problems. The experimental results were used as reference values to evaluate the accuracy of the developed numerical models.

6. Conclusions and Future Activities

For airframe structural components, it is crucial to quickly estimate the severity of the dynamic load conditions at which vibration loads could occur. The dynamic load increment plays a key role in affecting the structural strength assessment when the system’s natural frequencies are close to the excitation source characteristics. An engineering method of performing preliminary conservative evaluations of flight equipment is provided. Based on the resonance frequencies and participating masses, this paper analyzes the dynamic characteristics of an MLG conceived for high-speed applications against the operative vibration spectra. The mode shapes are sorted according to their participating mass entities, and less dominant modes are discarded to streamline the model. The FE-based results and those of the trial test are close, proving that the method can ensure good accuracy within the calculation and analysis process. Finally, the structural reliability of the MLG and its retraction actuator assembly are verified by a mechanical vibration test. Future activities could include the planning of an addressed modal analysis useful for the validation of the actual dynamic properties. The comparison of the present results with measurements carried out at the helicopter level would be another interesting task that would allow for the investigation of the influence of the actual boundary conditions with greater detail.

Author Contributions

Conceptualization, all; methodology, A.C., M.A. and C.C.T.; software, M.A. and C.C.T.; validation, all; formal analysis, A.C., M.A. and C.C.T.; investigation, all; resources, M.B.; data curation, M.A. and C.C.T.; writing—original draft preparation, M.A. and C.C.T.; writing—review and editing, M.A. and C.C.T.; supervision, A.C., M.A., C.C.T., V.Q., M.B. and M.C.; project administration, A.C.; funding acquisition, M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Clean Sky 2 Joint Undertaking under the European Union’s Horizon 2020 research and innovation program under grant agreement No. CS2-GAM-FRC-2014-2015 and the following extensions.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors are grateful to the four reviewers for their constructive comments to improve the paper. The authors thank, in particular, Julie Ganivet and Jaroslaw Zakrzewski from Airbus Helicopter for providing the technical requirements and the follow-up during the development and qualification. The authors would also like to sincerely thank all the people who contributed to the project’s activities and, in particular, the Magnaghi Aeronautica Engineering and Lab Test Team.

Conflicts of Interest

The authors declare no conflict of interest. Maurizio Arena and Martina Castaldo are employees of Magnaghi Aeronautica of MA Group Company. The paper reflects the views of the scientists, and not the company.

References

  1. Kehoe, M.W. Aircraft Ground Vibration Testing at NASA Ames–Dryden Flight Research Facility; NASA TM–88272; National Aeronautics and Space Administration, AMES Research Center, Dryden Flight Research Facility: Edwards, CA, USA, 1987. [Google Scholar]
  2. Kehoe, M.W.; Voracek, D.F. Ground Vibration Test Results of a JetStar Airplane Using Impulsive Sine Excitation; NASA TM–100448; National Aeronautics and Space Administration, AMES Research Center, Dryden Flight Research Facility: Edwards, CA, USA, 1994. [Google Scholar]
  3. Cox, T.H.; Gilyard, G.B. Ground Vibration Test Results for Drones for Aerodynamic and Structural Testing (DAST)/Aeroelastic Research Wing (ARW–1R) Aircraft; NASA TM–85906; National Aeronautics and Space Administration, AMES Research Center, Dryden Flight Research Facility: Edwards, CA, USA, 1986. [Google Scholar]
  4. Afonso, F.; Vale, J.; Oliveira, É.; Lau, F.; Suleman, A. A review on non-linear aeroelasticity of high aspect-ratio wings. Prog. Aerosp. Sci. 2017, 89, 40–57. [Google Scholar] [CrossRef]
  5. Gao, Y.; An, G.; Zhi, C. Chapter 9–Airworthiness verification test of the flight control system. In Test Techniques for Flight Control Systems of Large Transport Aircraft; Gao, Y., An, G., Zhi, C., Eds.; Academic Press: Cambridge, MA, USA, 2021; pp. 631–672. [Google Scholar]
  6. Chang, C.; Hodges, D. Modeling of Ground Vibration Testing for Highly Flexible Aircraft. In Proceedings of the 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Newport, RI, USA, 1–4 May 2006. [Google Scholar]
  7. Göge, D.; Böswald, M.; Füllekrug, U.; Lubrina, P. Ground vibration testing of large aircraft-state-of-the-art and future perspectives. In Proceedings of the 25th International Modal Analysis Conference (IMAC XXV), Orlando, FL, USA, 19–22 January 2007. [Google Scholar]
  8. Füllekrug, U.; Gloth, G.; Sinapius, M.; Fargette, P.; Levadoux, B.; Lubina, P.; Schaak, H. Tasks for improvement in ground vibration testing of large aircraft. In Proceedings of the International Forum on Aeroelasticity and Structural Dynamics (IFASD 2001), Madrid, Spain, 5–7 June 2001. [Google Scholar]
  9. Pickrel, C.R.; Foss, G.C.; Phillips, A.; Allemang, R.J.; Brown, D.L. New concepts GVT. In Proceedings of the 24th International Modal Analysis Conference (IMAC XXIV), St. Louis, MO, USA, 30 January–2 February 2006. [Google Scholar]
  10. Lubrina, P. Ground vibration experiments on large civil aircraft for engine imbalance purpose. In Proceedings of the International Forum on Aeroelasticity and Structural Dynamics (IFASD 2003), Amsterdam, The Netherlands, 4–6 June 2003. [Google Scholar]
  11. Gloth, G.; Degener, M.; Füllekrug, U.; Gschwilm, J.; Sinapius, M.; Fargette, P.; Levadoux, B.; Lubrina, P. Experimental investigation of new GVT concepts for large aircraft. Sound Vib. 2001, 35, 14–18. [Google Scholar]
  12. Simsiriwong, J.; Sullivan, R.W. Experimental Vibration Analysis of a Composite UAV Wing. Mech. Adv. Mater. Struct. 2012, 19, 196–206. [Google Scholar] [CrossRef]
  13. Sullivan, R.; Rais-Rohani, M.; Lacy, T.; Alday, N. Structural testing of an ultralight UAV composite wing. In Proceedings of the 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, Newport, RI, USA, 1–4 May 2006. [Google Scholar]
  14. Sullivan, R.W.; Hwang, Y.; Rais-Rohani, M.; Lacy, T. Structure analysis and testing of an ultralight UAV carbon composite wing. J. Aircr. 2009, 46, 814–820. [Google Scholar] [CrossRef]
  15. Simsiriwong, J.; Sullivan, R. Vibration testing of a carbon composite fuselage. In Proceedings of the 23rd American Society for Composites Conference, Memphis, TN, USA, 9–11 September 2008. [Google Scholar]
  16. Gupta, A.; Seiler, P.; Danowsky, B. Ground vibration tests on a flexible flying wing aircraft. In Proceedings of the AIAA Atmospheric Flight Mechanics Conference, AIAA SciTech, San Diego, CA, USA, 4–8 January 2016. AIAA Paper 2016-1753. [Google Scholar]
  17. Kerschen, G.; Peeters, M.; Golinval, J.C.; Stéphan, C. Nonlinear Modal Analysis of a Full-Scale Aircraft. J. Aircr. 2013, 50, 1409–1419. [Google Scholar] [CrossRef]
  18. Ahlquist, J.; Carreño, J.; Climent, H.; de Diego, R.; de Alba, J. Assessment of nonlinear structural response in A400M GVT. In Proceedings of the 28th International Modal Analysis Conference (IMAC XXVIII), Jacksonville, FL, USA, 1–4 February 2010. [Google Scholar]
  19. Goege, D. Fast Identification and Characterization of Nonlinearities in Experimental Modal Analysis of Large Aircraft. J. Aircr. 2007, 44, 399–409. [Google Scholar] [CrossRef]
  20. Pritchard, J. Overview of Landing Gear Dynamics. J. Aircr. 2001, 38, 130–137. [Google Scholar] [CrossRef]
  21. Somieski, G. Shimmy analysis of a simple aircraft nose landing gear model using different mathematical methods. Aerosp. Sci. Technol. 1997, 1, 545–555. [Google Scholar] [CrossRef]
  22. Esmailzadeh, E.; Farzaneh, K. Shimmy Vibration Analysis of Aircraft Landing Gears. J. Vib. Control 1999, 5, 45–56. [Google Scholar] [CrossRef]
  23. Feng, F.; Nie, H.; Zhang, M.; Peng, Y. Effect of Torsional Damping on Aircraft Nose Landing-Gear Shimmy. J. Aircr. 2015, 52, 561–568. [Google Scholar] [CrossRef]
  24. Rahmani, M.; Behdinan, K. Interaction of torque link freeplay and Coulomb friction nonlinearities in nose landing gear shimmy scenarios. Int. J. Non-Linear Mech. 2020, 119, 103338. [Google Scholar] [CrossRef]
  25. Du, X.; Xu, Y.; Liu, Q.; Liu, C.; Yue, X.; Liu, X.; Kurths, J. Shimmy dynamics in a dual-wheel nose landing gear with freeplay under stochastic wind disturbances. Nonlinear Dyn. 2024, 112, 2477–2499. [Google Scholar] [CrossRef]
  26. Li, Y.; Howcroft, C.; Neild, S.A.; Jiang, J.Z. Using Continuation Analysis to Identify Shimmy-Suppression Devices for an Aircraft Main Landing Gear. J. Sound Vib. 2017, 408, 234–251. [Google Scholar] [CrossRef]
  27. Orlando, C.; Alaimo, A. A Robust Active Control System for Shimmy Damping in the Presence of Free Play and Uncertainties. Mech. Syst. Signal Process. 2017, 84, 551–569. [Google Scholar] [CrossRef]
  28. Rahmani, M.; Behdinan, K. Parametric study of a novel nose landing gear shimmy damper concept. J. Sound Vib. 2019, 457, 299–313. [Google Scholar] [CrossRef]
  29. Rahmani, M.; Behdinan, K. On the effectiveness of shimmy dampers in stabilizing nose landing gears. Aerosp. Sci. Technol. 2019, 91, 272–286. [Google Scholar] [CrossRef]
  30. Mustashin, M.S.; Rahmani, M.; Behdinan, K. Experimental characterization of a novel nose landing gear shimmy damper using a small-scale test rig. Aerosp. Sci. Technol. 2021, 112, 106625. [Google Scholar] [CrossRef]
  31. Li, F.; Zhao, Y. Amplitude reduction optimization of time delay semi-active control for aircraft landing gear shimmy. J. Phys. Conf. Ser. 2022, 2230, 012009. [Google Scholar] [CrossRef]
  32. Rahmani, M.; Behdinan, K. Structural design and optimization of a novel shimmy damper for nose landing gears. Struct. Multidiscip. Optim. 2020, 62, 2783–2803. [Google Scholar] [CrossRef]
  33. Biehl, F.A. Aircraft Landing Gear Brake Squeal and Strut Chatter Investigation. In The Shock and Vibration Bulletin; Naval Research Laboratory: Washington, DC, USA, 1969. [Google Scholar]
  34. Travis, M.H. Predicting Landing Gear Carbon Brake Vibration and Performance via Subscale Test and Analysis; SAE Technical Paper 2005-01-3437; SAE International: Warrendale, PA, USA, 2005. [Google Scholar] [CrossRef]
  35. Yin, Q.; Jiang, J.Z.; Neild, S.A.; Nie, H. Investigation of gear walk suppression while maintaining braking performance in a main landing gear. Aerosp. Sci. Technol. 2019, 91, 122–135. [Google Scholar] [CrossRef]
  36. Jiao, Z.; Bai, N.; Liu, J.; Li, Y.; Wang, Z.; Sun, D.; Qi, P.; Shang, Y. Aircraft anti-skid braking control technology: A review. Acta Aeronaut. Astronaut. Sin. 2022, 43, 1–27. [Google Scholar]
  37. Sinou, J.-J.; Dereure, O.; Mazet, G.-B.; Thouverez, F.; Jezequel, L. Friction-induced vibration for an aircraft brake system—Part 1: Experimental approach and stability analysis. Int. J. Mech. Sci. 2006, 48, 536–554. [Google Scholar] [CrossRef]
  38. Arena, M.; Chiariello, A.; Castaldo, M.; Di Palma, L. Vibration Response Aspects of a Main Landing Gear Composite Door Designed for High-Speed Rotorcraft. Aerospace 2021, 8, 52. [Google Scholar] [CrossRef]
  39. Liu, R.; Li, Z.; Xu, W.; Yang, X.; Zhang, D.; Yao, Z.; Yang, K. Dynamic Analysis and Experiment of a Space Mirror Based on a Linear State Space Expression. Appl. Sci. 2021, 11, 5379. [Google Scholar] [CrossRef]
  40. Wang, L.; Zhong, H. A time finite element method for structural dynamics. Appl. Math. Model. 2017, 41, 445–461. [Google Scholar] [CrossRef]
  41. MSC Nastran®. Quick Reference Guide 2021; MSC Software: Newport Beach, CA, USA, 2021. [Google Scholar]
  42. MMPDS-16; Metallic Materials Properties Development and Standardization. Battelle Memorial Institute: Columbus, OH, USA, 2021.
  43. Wijker, J. Miles’ Equation in Random Vibrations; Solid Mechanics and Its Applications; Springer: Cham, Switzerland, 2018; Volume 248. [Google Scholar] [CrossRef]
  44. Vince, A.; Abraham, A. Building Better Products with Finite Element Analysis; Onward Press: Santa Fe, NM, USA, 1999. [Google Scholar]
  45. Budynas, R.; Nisbett, K. Shigley’s Mechanical Engineering Design; McGraw-Hill: New York, NY, USA, 2008. [Google Scholar]
  46. Lee, Y.-L. Stress-Based Fatigue Analysis and Design. In Fatigue Testing and Analysis; Elsevier: Oxford, UK, 2005. [Google Scholar]
  47. RACER Compound Helicopter. Clean Aviation. Available online: https://www.clean-aviation.eu/racer-compound-helicopter (accessed on 1 December 2023).
  48. Clean Sky 2–ANGELA. Available online: https://www.maregroup.it/cleansky2/portfolio/angela/ (accessed on 1 December 2023).
  49. CIRA. ANGELA Project, Vibrational Qualification Campaign of the RACER Compound Helicopter Landing Gears Completed. Available online: https://www.cira.it/en/competences/strutture-e-materiali/progetto-angela-superati-i-test-di-qualifica-vibrazionale-per-i-carrelli-d’atterraggio-del-racer-compound-helicopter (accessed on 1 December 2023).
  50. RTCA-DO-160-F; Environmental Conditions and Test Procedures for Airborne Equipment. RTCA, Inc.: Washington, DC, USA, 2007.
  51. Shaker Fixture Design. Vibration Research. Available online: https://vibrationresearch.com/blog/shaker-fixtures-and-fixture-design (accessed on 1 December 2023).
  52. ECSS-E-ST-10-03C; Space Engineering: Testing. ECSS: Noordwijk, The Netherlands, 2012.
  53. Orefice, M.; Bruno, G.; Tartaglia, C.C.; Di Caprio, F.; Menichino, A.; Quaranta, V.; Belardo, M. Vibration test campaign on ALM flanges of bladder fuel tanks for Tiltrotor application. In Proceedings of the 2022 IEEE 9th International Workshop on Metrology for AeroSpace (MetroAeroSpace), Pisa, Italy, 27–29 June 2022; pp. 38–42. [Google Scholar]
Figure 1. MLG system: (a) leg with the actuation system; (b) extended and retracted attitudes.
Figure 1. MLG system: (a) leg with the actuation system; (b) extended and retracted attitudes.
Aerospace 11 00130 g001
Figure 2. Main landing gear FE models: (a) actuator; (b) leg components.
Figure 2. Main landing gear FE models: (a) actuator; (b) leg components.
Aerospace 11 00130 g002
Figure 3. Boundary condition details: actuator and MLG leg: (a) MLG leg; (b) side-brace actuator.
Figure 3. Boundary condition details: actuator and MLG leg: (a) MLG leg; (b) side-brace actuator.
Aerospace 11 00130 g003
Figure 4. FEM contact details: (a) bearing interface inside S/A chamber; (b) expanded view of articulated links; (c) sealed interface regions of the actuator.
Figure 4. FEM contact details: (a) bearing interface inside S/A chamber; (b) expanded view of articulated links; (c) sealed interface regions of the actuator.
Aerospace 11 00130 g004
Figure 5. MLG random load spectrum (performance Grms = 2.75 g; endurance Grms = 3.89 g).
Figure 5. MLG random load spectrum (performance Grms = 2.75 g; endurance Grms = 3.89 g).
Aerospace 11 00130 g005
Figure 6. FE equivalent steady analysis flowchart.
Figure 6. FE equivalent steady analysis flowchart.
Aerospace 11 00130 g006
Figure 7. MLG normal modes: MLG leg (a) first and second flexural mode f = 33.2 Hz, f = 34.8 Hz; retraction actuator (b) first flexural mode, f = 35.5 Hz.
Figure 7. MLG normal modes: MLG leg (a) first and second flexural mode f = 33.2 Hz, f = 34.8 Hz; retraction actuator (b) first flexural mode, f = 35.5 Hz.
Aerospace 11 00130 g007
Figure 8. MLG trial test. Setup configuration for excitation in the (a) X direction; (b) Z direction.
Figure 8. MLG trial test. Setup configuration for excitation in the (a) X direction; (b) Z direction.
Aerospace 11 00130 g008
Figure 9. MLG trial test. Setup configuration for excitation in the Y direction.
Figure 9. MLG trial test. Setup configuration for excitation in the Y direction.
Aerospace 11 00130 g009
Figure 10. MLG measuring point on the wheel.
Figure 10. MLG measuring point on the wheel.
Aerospace 11 00130 g010
Figure 11. SBA extended measuring point in the middle of the actuator.
Figure 11. SBA extended measuring point in the middle of the actuator.
Aerospace 11 00130 g011
Figure 12. SBA extended trial test. Setup configuration for excitation in the X (a), Y (b), and Z (c) direction.
Figure 12. SBA extended trial test. Setup configuration for excitation in the X (a), Y (b), and Z (c) direction.
Aerospace 11 00130 g012
Figure 13. MLG deflections at equivalent static loads: MLG leg (a) first and second flexural mode; retraction actuator (b) first flexural mode.
Figure 13. MLG deflections at equivalent static loads: MLG leg (a) first and second flexural mode; retraction actuator (b) first flexural mode.
Aerospace 11 00130 g013
Figure 14. Deformation pattern with the Von Mises stresses: MLG leg (a); retraction actuator (b).
Figure 14. Deformation pattern with the Von Mises stresses: MLG leg (a); retraction actuator (b).
Aerospace 11 00130 g014
Figure 15. MLG and actuator stress results on most loaded parts. (a) Main fitting (fty = 400 MPa), (b) sliding rod and wheel axle (fty = 1586 MPa), (c) torque links (fty = 792 MPa), (d) piston rod (fty = 1000 MPa) and end fitting (fty = fcy = 806 MPa).
Figure 15. MLG and actuator stress results on most loaded parts. (a) Main fitting (fty = 400 MPa), (b) sliding rod and wheel axle (fty = 1586 MPa), (c) torque links (fty = 792 MPa), (d) piston rod (fty = 1000 MPa) and end fitting (fty = fcy = 806 MPa).
Aerospace 11 00130 g015aAerospace 11 00130 g015bAerospace 11 00130 g015c
Figure 16. MLG coupled with fixture.
Figure 16. MLG coupled with fixture.
Aerospace 11 00130 g016
Figure 17. (a) SBA retracted and (b) extended coupled with fixture.
Figure 17. (a) SBA retracted and (b) extended coupled with fixture.
Aerospace 11 00130 g017
Figure 18. Racer H/C with vibration zones.
Figure 18. Racer H/C with vibration zones.
Aerospace 11 00130 g018
Figure 19. Effect of gravity on the deformation of the MLG system in (a) −Z direction; (b) +Y direction.
Figure 19. Effect of gravity on the deformation of the MLG system in (a) −Z direction; (b) +Y direction.
Aerospace 11 00130 g019
Figure 20. MLG (a) and SBA (b) reference system.
Figure 20. MLG (a) and SBA (b) reference system.
Aerospace 11 00130 g020
Figure 21. MLG location of the control accelerometers.
Figure 21. MLG location of the control accelerometers.
Aerospace 11 00130 g021
Figure 22. MLG location of the measurement accelerometers.
Figure 22. MLG location of the measurement accelerometers.
Aerospace 11 00130 g022
Figure 23. MLG setup configuration for excitation in the Z direction.
Figure 23. MLG setup configuration for excitation in the Z direction.
Aerospace 11 00130 g023
Figure 24. SBA extended: location of the control and measurement accelerometers.
Figure 24. SBA extended: location of the control and measurement accelerometers.
Aerospace 11 00130 g024
Figure 25. SBA extended: location of the accelerometers AC1 and AC3.
Figure 25. SBA extended: location of the accelerometers AC1 and AC3.
Aerospace 11 00130 g025
Figure 26. SBA extended: setup configuration for excitation in the X direction.
Figure 26. SBA extended: setup configuration for excitation in the X direction.
Aerospace 11 00130 g026
Figure 27. MLG: FRF AC1-Z/AVG_CTR before (blue) and after (red) vibration tests.
Figure 27. MLG: FRF AC1-Z/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g027
Figure 28. MLG: FRF AC2-Z/AVG_CTR before (blue) and after (red) vibration tests.
Figure 28. MLG: FRF AC2-Z/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g028
Figure 29. MLG: FRF AC3-Z/AVG_CTR before (blue) and after (red) vibration tests.
Figure 29. MLG: FRF AC3-Z/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g029
Figure 30. MLG: FRF AC4-Z/AVG_CTR before (blue) and after (red) vibration tests.
Figure 30. MLG: FRF AC4-Z/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g030
Figure 31. SBA in ext. conf.: FRF AC1-X/AVG_CTR before (blue) and after (red) vibration tests.
Figure 31. SBA in ext. conf.: FRF AC1-X/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g031
Figure 32. SBA in ext. conf.: FRF AC2-X/AVG_CTR before (blue) and after (red) vibration tests.
Figure 32. SBA in ext. conf.: FRF AC2-X/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g032
Figure 33. SBA in ext. conf.: FRF AC3-X/AVG_CTR before (blue) and after (red) vibration tests.
Figure 33. SBA in ext. conf.: FRF AC3-X/AVG_CTR before (blue) and after (red) vibration tests.
Aerospace 11 00130 g033
Figure 34. SBA in ext. conf.: FRF AC4-X/AVG_CTR before (blue) and after (red) vibration tests
Figure 34. SBA in ext. conf.: FRF AC4-X/AVG_CTR before (blue) and after (red) vibration tests
Aerospace 11 00130 g034
Table 1. MLG system: material properties of main items [42].
Table 1. MLG system: material properties of main items [42].
ItemComponentMaterialSpecification
[-][-][-][-]
MLG LegMain FittingAl 7050-T7452AMS 4108
Sliding Rod—Wheel AxleSteel 300MAMS 6257
Torque LinksTi-6Al-4VAMS 4928
ActuatorEnd Cap—End FittingTi-6Al-4VAMS 4928
Cylinder—Piston RodSteel 15-5PHAMS 5659
Table 2. Three-dimensional FE model’s total number of nodes and elements.
Table 2. Three-dimensional FE model’s total number of nodes and elements.
FE EntityNumber
Nodes1,384,642
3D elements, ctet46,270,015
1D elements, cgap contact32,785
0D elements, conm21
Rigid elements, rbe36
Table 3. W&B mass properties.
Table 3. W&B mass properties.
Mass, m19.5[kg]Aerospace 11 00130 i001
Inertia moment, xy
Ixx = Iyy
0.229[kgm2]
Inertia moment, z
Izz
0.331[kgm2]
Table 4. Helicopter sine tones and random spectrum.
Table 4. Helicopter sine tones and random spectrum.
Frequency BandAmplitudeDescription
[Hz][g, linear], [g2/Hz, Random][-]
23.3±1.05/rev MR
29.2±1.01/rev LR
46.5±0.6710/rev MR
175.2±0.676/rev LR
10–20000.01 (10; 300 Hz), 0.001 (f ≤ 2000) performance Grms = 2.75 g
0.02 (10; 300 Hz), 0.002 (f ≤ 2000) endurance Grms = 3.89 g
linear acceleration confidential (indicated as normalized values); 1/rev: rotor vibration cycle occurring once for every complete rotor revolution.
Table 5. MLG dynamic parameters.
Table 5. MLG dynamic parameters.
FrequencyModal MassDescription
[Hz][kg][-]
33.241.1MLG leg first bending
34.827.8MLG leg second bending
35.58.0Actuator first bending
Table 6. Trial test. Comparison between numerical and experimental results.
Table 6. Trial test. Comparison between numerical and experimental results.
AxisFirst Numerical Frequency [Hz]First Experimental Frequency [Hz]Numerical Q FactorExperimental Q Factor
X33.227.24201.86
Y30.2810.84
Z31.2311.45
Table 7. SBA extended trial test. Comparison between numerical and experimental results.
Table 7. SBA extended trial test. Comparison between numerical and experimental results.
AxisFirst Numerical Frequency [Hz]First Experimental Frequency [Hz]Numerical Q FactorExperimental Q Factor
X35.5No resonance up to 200 Hz20n/a
Y29.037.03
Z27.393.31
Table 8. MLG system: peak stress summary for worst condition (endurance levels).
Table 8. MLG system: peak stress summary for worst condition (endurance levels).
AssyComponentStress ComponentYield StrengthLoad Case
[-][-][MPa][MPa][-]
MLG legMain Fitting139.0 (Max P)400.0x direction
Sliding Rod603.7 (Max P)1586.0x direction
Wheel Axle216.9 (Max P)1586.0z direction
Upper Torque Link168.7 (Max P)792.0x direction
Lower Torque Link176.1 (Max P)792.0x direction
ActuatorEnd Cap<40.0806.0x direction
(SBA reference frame)
End Fitting(−)43.6 (Max P)806.0
Cylinder<40.01000.0
Piston Rod144.3 (Max P)1000.0
Table 9. MLG system: dwell testing duration and number of cycles.
Table 9. MLG system: dwell testing duration and number of cycles.
TestDurationCycles Required, ni
[-][min][–]
MLGActuator
Performance10.019,920 21,283
Endurance20.039,840 42,566
Performance10.019,920 21,283
Table 10. MLG system: preliminary fatigue damage for main items.
Table 10. MLG system: preliminary fatigue damage for main items.
AssyComponentMax Principal
(Performance/Endurance)
Cycles to Failure,
Ni
Damage,
D
[-][-][MPa][-] i n i N f i
MLG legMain Fitting(39.4; 55.5)1.0 × 1070.000
Sliding Rod(184.1; 259.6)1.0 × 1060.415
Wheel Axle(45.0; 63.4)1.0 × 1070.000
Upper Torque Link(47.7; 67.3)1.0 × 1070.000
Lower Torque Link(49.8; 70.3)1.0 × 1070.000
ActuatorEnd Cap<40 MPa1.0 × 1070.000
End Fitting<40 MPa1.0 × 1070.000
Cylinder<40 MPa1.0 × 1070.000
Piston Rod(118.1; 144.0)1.0 × 1070.004
Table 11. Equipment under test.
Table 11. Equipment under test.
DescriptionSerial NumberLength [mm]EUT Mass [kg]
MLGTS100-G1033-001065.462.0
SBATS100-A1051-00EXTENDED: 1326.7
RETRACTED: 810.8
14.0
Table 12. Vibration test levels.
Table 12. Vibration test levels.
Sine-on-Random
PERFORMANCE (600 s)
Sine-on-Random
ENDURANCE (1200 s)
Background RandomBackground Random
Frequency [Hz]ASD [g2/Hz]Frequency [Hz]ASD [g2/Hz]
100.01100.02
3000.013000.02
20000.00120000.002
Sweeping Sine TonesSweeping Sine Tones
ItemFrequency [Hz]23.2529.2046.50175.2023.2529.2046.50175.20
MLGX-Axis [g] *1.000.670.670.671.000.670.670.67
Y-Axis [g] *1.000.670.670.671.000.670.670.67
Z-Axis [g] *1.000.190.190.191.000.330.330.33
SBA
Retracted
X-Axis [g] *1.000.670.670.671.000.670.670.67
Y-Axis [g] *1.000.670.670.671.000.670.670.67
Z-Axis [g] *1.000.190.190.191.000.330.330.33
SBA
Extended
X-Axis [g] *1.000.670.670.671.000.670.670.67
Y-Axis [g] *0.880.751.001.001.001.001.001.00
Z-Axis [g] *0.880.751.001.001.001.001.001.00
* normalized to the maximum value.
Table 13. MLG: shift in the response of AC4 before and after the test along the Z axis direction.
Table 13. MLG: shift in the response of AC4 before and after the test along the Z axis direction.
Resonance IDFrequency Shift [%]Amplitude Shift [%]
10.290.00
2−1.112.54
30.83−3.96
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chiariello, A.; Carandente Tartaglia, C.; Arena, M.; Quaranta, V.; Bruno, G.; Belardo, M.; Castaldo, M. Vibration Qualification Campaign on Main Landing Gear System for High-Speed Compound Helicopter. Aerospace 2024, 11, 130. https://doi.org/10.3390/aerospace11020130

AMA Style

Chiariello A, Carandente Tartaglia C, Arena M, Quaranta V, Bruno G, Belardo M, Castaldo M. Vibration Qualification Campaign on Main Landing Gear System for High-Speed Compound Helicopter. Aerospace. 2024; 11(2):130. https://doi.org/10.3390/aerospace11020130

Chicago/Turabian Style

Chiariello, Antonio, Carmine Carandente Tartaglia, Maurizio Arena, Vincenzo Quaranta, Giovanni Bruno, Marika Belardo, and Martina Castaldo. 2024. "Vibration Qualification Campaign on Main Landing Gear System for High-Speed Compound Helicopter" Aerospace 11, no. 2: 130. https://doi.org/10.3390/aerospace11020130

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop