Journal of Sound and Vibration xxx (2018) 1e89
Contents lists available at ScienceDirect
Journal of Sound and Vibration journal homepage: www.elsevier.com/locate/jsvi
Literature review of models on tire-pavement interaction noise Tan Li*, Ricardo Burdisso, Corina Sandu Department of Mechanical Engineering, Virginia Tech, Blacksburg, VA 24061, USA
a r t i c l e i n f o
a b s t r a c t
Article history: Received 29 April 2017 Received in revised form 18 December 2017 Accepted 10 January 2018 Available online xxx
Tire-pavement interaction noise (TPIN) becomes dominant at speeds above 40 km/h for passenger vehicles and 70 km/h for trucks. Several models have been developed to describe and predict the TPIN. However, these models do not fully reveal the physical mechanisms or predict TPIN accurately. It is well known that all the models have both strengths and weaknesses, and different models fit different investigation purposes or conditions. The numerous papers that present these models are widely scattered among thousands of journals, and it is difficult to get the complete picture of the status of research in this area. This review article aims at presenting the history and current state of TPIN models systematically, making it easier to identify and distribute the key knowledge and opinions, and providing insight into the future research trend in this field. In this work, over 2000 references related to TPIN were collected, and 74 models were reviewed from nearly 200 selected references; these were categorized into deterministic models (37), statistical models (18), and hybrid models (19). The sections explaining the models are self-contained with key principles, equations, and illustrations included. The deterministic models were divided into three sub-categories: conventional physics models, finite element and boundary element models, and computational fluid dynamics models; the statistical models were divided into three sub-categories: traditional regression models, principal component analysis models, and fuzzy curve-fitting models; the hybrid models were divided into three sub-categories: tire-pavement interface models, mechanism separation models, and noise propagation models. At the end of each category of models, a summary table is presented to compare these models with the key information extracted. Readers may refer to these tables to find models of their interest. The strengths and weaknesses of the models in different categories were then analyzed. Finally, the modeling trend and future direction in this area are given. © 2018 Elsevier Ltd. All rights reserved.
Keywords: Tire noise Deterministic model Statistical model Hybrid model
1. Introduction The term tire-pavement interaction noise was first found in the literature in 1976 (Thrasher et al., 1976 [1] and Miller and Thrasher, 1976 [2]), and later in 1986 (Pottinger et al., 1986 [3]) and 1998 (Donavan et al., 1998 [4]). The acronym TPIN (for tirepavement interaction noise) was first formed in 2008 (Khazanovich and Izevbekhai, 2008 [5]), and later found in 2013 (Syamkumar et al., 2013 [6]). Tire-pavement interaction noise is also known as tire-road interaction noise, tire/pavement
* Corresponding author. E-mail addresses:
[email protected] (T. Li),
[email protected] (R. Burdisso),
[email protected] (C. Sandu). https://doi.org/10.1016/j.jsv.2018.01.026 0022-460X/© 2018 Elsevier Ltd. All rights reserved.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
2
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
noise, tire/road noise (TRN), or tire noise. Tire-pavement interaction noise is defined as the noise emitted from a rolling tire as a result of the interaction between the tire and the road surface (Sandberg and Ejsmont, 2002 [7]). Modern research on TPIN started around mid-1970's as engines became less noisy and the aerodynamic design of the vehicle body was optimized. Sandberg and Ejsmont (1995) [8] have compiled a computerized bibliography termed “TRN Bibliography” totaling 1280 documents as of August 1994. However, most of these references were from special conference proceedings and industrial/government reports, and not from archival journals. To date, a larger number of references can be found (totaling 2654 published during the period 1970e2015, as displayed in Fig. 1), which have greatly advanced the modeling of tirepavement interaction noise. However, TPIN is still not fully understood, and it is difficult to describe or predict TPIN accurately; there are large differences in various findings, probably because of the many complex generation mechanisms involved. Although a tire seems to be a simple structure at the first glance, the combination of its geometrical properties and multi-layer composite structure leads to a highly complex vibrational behavior (Sabiniarz and Kropp, 2010 [9]). Tire-pavement interaction is the dominant source of noise at speeds above 40 km/h for passenger vehicles and 70 km/h for trucks. There are several reasons for reducing tire-pavement interaction noise in different countries: to pass the legal approval for exterior noise level; to reduce the annoyance level in the interior space of the car and increase passenger comfort (Harrison, 2004 [10]); to avoid the import of foreign tires (Chen, 2014 [11]). It is certain that accurate and reliable models on TPIN are desired by vehicle manufacturers and tire companies, as these models will be able to reduce the development time (Mohamed et al., 2013 [12]). Several models have been developed to describe and predict tire vibrations and TPIN. However, all these models have restricted applications in some sense (Dare, 2012 [13]). For example, some models describe only one or two generation mechanisms; some assume two-dimensional tire and pavement; some include only a limited number of tire/pavement combinations, such as a slick tire on a rough road; some calculate the noise level at the source without considering sound propagation; some are concerned only with the structural dynamics of the tire, and neglect the effect of the nonlinear aerodynamics surrounding the tire. It is difficult to evaluate the accuracy of a specific model because of the different objectives and lack of validation (Kuijpers and van Blokland, 2001 [14]). In addition, most of the current models focus only on passenger car tires. Very few studies have reported noise models for truck tires that might have more influence on traffic noise. Because of the increased popularity of porous asphalt pavement, a large number of models have included pavement absorption and mechanical impedance as the parameters of interest. In general, there are three acoustic modeling approaches: wave, energy, and ray acoustics (Crocker, 2007 [15]). The wave acoustics approach, commonly adopted by finite element software, is used for a small enclosed space, especially if its shape is simple and has well-defined boundary conditions. The energy acoustics approach is often used as a statistical approach for measuring sound intensity. The ray acoustics approach, together with the multiple-image-source concept for reflection calculations, can be used to investigate sound propagation in the atmosphere. Most of the current TPIN models have used one or two approaches mentioned above. In this literature review work, the present authors collected over 2000 references related to tire noise, published during the period from 1960's to 2010's (Li, 2017 [16]). The present authors cannot guarantee that all the references have been covered, but a vast majority has been included. In the end, 74 models were reviewed from nearly 200 selected references. The models were divided into three categories based on the characteristics of the approach used: deterministic model, statistical model, and hybrid model. A brief introduction of the three categories is presented in Table 1. The deterministic models normally focused on low frequencies (below 500 Hz). However, the main spectral content of the actual tire noise is located in the range of 600e1200 Hz. Therefore, these models tend to have considerable errors. The statistical and hybrid models normally cover the entire dominant frequency of tire noise, resulting in better accuracy. However, most of the models investigate only the pavement parameters. Both advantages and disadvantages exist for all the three categories of TPIN models. In fact, one specific model generally focuses only on a few points of interest (physics or
Fig. 1. Count of records in literature (generated via Engineering Village on August 12, 2016, using search query for “tire” and “noise” in subject/title/abstract from the databases of Compendex, Inspec, and NTIS).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
3
Table 1 Introduction of three categories of TPIN models. Deterministic
Statistical
Hybrid
Method
Physics without experimental data
Example
Brinkmeier et al., 2008 [17]: structurally and aerodynamically radiated noise components Low frequency only, computationally intensive, 17 dB error
Correlation of measured noise data and tire-pavement parameters Khazanovich et al., 2008 [5]: different pavements and same tire with curve fitting
Incorporating physical principles into empirical model Domenichini et al., 1999 [18]: effect of pavement texture on TPIN Mostly pavement parameters only, without knowledge of tire inputs; intensive data for training Optimization of the combination of tire and pavement
Drawback
Major application [14]
Tire design (qualitative)
No knowledge of noise generation mechanism
Pavement design (quantitative)
parameters) owing to the complexity of the TPIN mechanism. For different research purposes and/or under different conditions, one model might have certain advantages over the others in the same category of models. The purpose of this paper is to present the existing models from literature, compare these models, state their strengths and weaknesses, and provide insight into future research direction in this area. The organization of this paper is as follows. In section 2, the deterministic models are reviewed; in section 3, the statistical models are reviewed; in section 4, the hybrid models are reviewed; in section 5, instructions for readers to use the models are given; in section 6, conclusions of this literature review are given. 2. Deterministic models 2.1. Introduction Deterministic models are also called analytical or numerical models. Recent methods, such as finite element method (FEM) (Lou, 2007 [19]) and boundary element method (BEM) (Wang et al., 2011 [20]; Nakajima et al., 1993 [21]), require large computational effort. A large number of parameters are also required, such as tire geometry, deformation of the tire, material parameters, and boundary conditions (Yang et al., 2013 [22]). On the other hand, it is possible to investigate the influence of a single parameter on TPIN using deterministic models (Biermann et al., 2007 [23]). In general, deterministic models need several material or construction parameters as inputs for tire modeling. For example, Larsson and Kropp model (1992) [24e26] requires 24 input parameters (6 geometrical parameters and 18 material parameters), some of which are presented below as an example in Table 2. Anderson et al., (2001) [27] also presented a measurement method for collecting these tire material data. In this section, 37 deterministic models were reviewed; they consist of 16 conventional physics models, 16 finite element and boundary element models, and 5 computational fluid dynamics (CFD) models. The model name, method, output noise parameter with the noise measurement technique used, noise generation mechanisms investigated, and input parameters of each model are listed and compared in tables as a summary in section 2.5. The summary tables can be used as a reference to direct the readers to the models of their interest. 2.2. Conventional physics models Conventional physics models typically use traditional analytical methods and classic mathematics, such as modal analysis, to present and investigate the dynamic behavior of the tire that is modeled as a ring, shell, or plate. Traditional acoustic equations are sometimes applied to analyze the structural wave propagation mode in the tire tread and wall, as well as the acoustical mode in the tire cavity. Sixteen models are reviewed in this subsection. In this paper, the heading for the corresponding model is written in this form: model name (authors, year); the models in each subsection are ordered chronologically. 2.2.1. Monopole (Hayden, 1971) Hayden (1971) [29] developed the first air pumping model, focusing on the noise generation at the leading edge of the tire contact patch where the tread is compressed. The model is based on the monopole theory, whose governing equation is given below.
pðr; uÞ ¼
ruQ 4pr
(1)
where p is the sound pressure (root mean square value), r is the distance between the receiver and the monopole, u is the angular frequency, r is the air density, and Q is the air volume change rate (root mean square value). The initial groove volume V0 is given by Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
4
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Table 2 Input tire parameters (slick 205/60R15) for Larsson and Kropp model (the data in brackets refer to loss factors for the corresponding parameters) (Modified from Andersson and Larsson, 2005 [28], Table 1; reprinted with permission from Ms. Brigitte Badelt of S. Hirzel Verlag on behalf of the European Acoustics Association). Category
Parameter
Value
Geometry
Length of tread plate Width of tread plate Thickness of tread layer Thickness of belt layer Density Young's modulus (real part) Poisson's ratio Density Young's modulus (real part) Poisson's ratio Radial bedding Circumferential bedding Lateral bedding Circumferential tension Lateral tension
2p 0.314 m 0.355 m 0.010 m 0.001 m 1175 kg/m3 25 MPa [0.05e0.70] 0.499 3720 kg/m3 580 GPa [0.02] 0.4 1.02 106 N/m3 [0.09] 1.02 106 N/m3 [0.09] 1.02 106 N/m3 [0.09] 42000 N/m2 [0.055] 42000 N/m2 [0.055]
Tread layer
Belt layer
V0 ¼ LDW
(2)
where L is the groove length in the circumferential direction, D is the groove depth, and W is the groove width. The air volume change rate Q is calculated using the equation
Q ðvÞ ¼
DV dV0 dLDW ¼ dDWv ¼ ¼ L=v Dt Dt
(3)
where v is the tire forward speed, DV is the volume change, Dt is the time duration, and d is the fractional change in the initial volume. The frequency u is determined by the repetition between consecutive groove cavities.
uðvÞ ¼ 2pv lp
(4)
where lp is the pitch distance between the groove cavities. Combining the above equations gives
pðr; vÞ ¼
rv2 dDW 2lp r
n
(5)
where n is the number of groove cavities along the tire width. 2.2.2. Unified (Plotkin and Stusnick, 1981) Plotkin and Stusnick (1981) [30] presented a unified set of analytical models to investigate the major noise generation mechanisms (impact and air pumping) systematically, considering the basic tire structural properties. Bohm's thin-shell equations (Bohm, 1967 [31]) were used to model the motion of a tire. The air pumping mechanism was analyzed based on the physical behavior (Plotkin et al., 1980 [32]). The influence of pavement texture was also included using Nilsson's linear excitation model (Nilsson, 1976 [33]). Then, sound radiation analysis was employed using Eberhardt's model (Eberhardt, 1979 [34]). The nomenclature for the model is listed in Table 3. 2.2.2.1. Thin shell equations for the motion of the tire. The radial tire belt was modeled as a uniform shell on a uniform elastic foundation. The two-dimensional shell equations for the tangential and radial displacements of the belt are as follows (Bohm, 1967 [31]).
EA 8 2 00 0 0 _ € _ _ ¼ pt ðf; tÞ v þ 2 m U ð v þ wÞ þ U ðv þ 2w vÞ 2 ðv00 þ w0 Þ þ kt v dt ðv0 U þ vÞ > > R0 > > < 00 EA T EI 0 2 0 € _ _ > w w þ 2 m U ð w vÞ þ U 2v ðw þ R Þ þ 2 ðv0 þ wÞ þ 4 wiv þ 2w00 þ w 02 ðw þ w00 Þ 0 > > R0 R0 R0 > : 0 _ ¼ pr ðf; tÞ þkr w þ dr ðw U þ wÞ
(6)
For a stationary tire, after setting all time derivatives to zero, the solutions of the above linear equations are Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
5
Table 3 Nomenclature for unified model. Denotation
Meaning
a A Ab Ac A(f0) An, Bn, Cn, Dn Cp d d r, d t jdetj E, F, G, H ER Es EA EI f(f1-f) F(f) G G(rjr0 ) GR h H(f1-f) kr, kt kA, kT l lb ls lv n N Nv p pi p0(t) p r, p t prs, prc, pts, ptc P Pavg Q r R R0 s s0 S t tb ts tt T0 U un V v, w wb wt wv W x, y, z yg ygi zs Dz
Speed of sound in air Tread area function Cross-sectional area of belt Cross-sectional area of cords in belt Function describing tread area Coefficients of Fourier series describing tire shape Phase velocity in belt Tread depth Radial and tangential damping, per unit length Magnitude of determinant Coefficients of inverted matrix Elastic modulus of rubber Elastic modulus of steel Extensional stiffness of tread and belts Bending stiffness of tread and belts; (EI)b is the value for the belt plies alone, and (EI)t is the value for the tread alone Function defining radial displacement at f due to unit force at f1 Force associated with tread-road interface Geometric amplification factor Green's function Shear modulus of rubber Distance from neutral surface of bending to center of tread Unit step function Radial and tangential bedding coefficients Wave number in air and tire Tread pitch length Tread block length An invariant to calculate bedding stiffness Tread void length Circumferential mode number Number of belt plies Number of voids Acoustic pressure Inflation pressure Sound pressure for a tread void with pitch l0, depth d0, and width w0 Radial and tangential external pressures Sine and cosine integrals of pr and pt; e.g., prs ¼ 1/p$!20pprsinnfdf Point force Average contact pressure Resonant Q factor; inverse of loss factor Vector from source to receiver; jrj ¼ r Local radius of curvature Unperturbed radius of tire Circumferential distance coordinate, measured from contact patch edge Coordinate on pavement Surface area Time Thickness of belt ply Sidewall thickness Tread thickness Tension in belt due to inflation pressure Speed of tire Normal velocity Tread void volume Tangential and radial displacement of belt Width of belt Effective width of tread Width of tread void Amplitude of harmonic component of radial displacement w, determined from the forced motion of the contact patch edge Cartesian coordinates with respect to contact patch center Shape of tread groove Shape of i-th groove in full tread pattern Pavement texture profile Average vertical compression of rubber Cord angle Matrix coefficients Dirac delta function Flow blockage factor Load distribution due to pavement texture Damping coefficient
a a, b, g, d, ε d(f-f1) εb
z h
(continued on next page)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
6
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Table 3 (continued ) Denotation
Meaning
q0 k k0 m r f fa, fb f0 Df Dfcp u U
Angle between sidewall and belt Local belt curvature ¼ l/R Curvature of unperturbed belt ¼ l/R0 Mass density (per unit length) of tread and belts Density of air Circumferential coordinate Coordinates of the leading edge and trailing edge of contact patch Circumferential coordinate fixed to tire; corresponds to f at t ¼ 0 Angle with respect to contact patch center Angle from center to edge of contact patch Angular frequency Angular velocity of tire Time derivative Spatial derivative; either v/vf or v/vs
(_) ()0
∞ 8 X > ðAn sin nf þ Bn cos nfÞ >
(7)
∞ X > > :w ¼ ðCn sin nf þ Dn cos nfÞ n¼0
where the coefficients are calculated in the following matrix format
2 3 E An 6 6 Bn 7 6 7 ¼ 1 6 F 4 Cn 5 jdetj 4 G H Dn 2
F E H G
G H I J
32 3 pts H 6 7 G7 76 ptc 7 J 54 prs 5 I prc
(8)
where the parameters in capital letters are given by
8 2 E ¼ a d þ ε2 dg2 > > > > > 2 > > F ¼ b d þ ε2 þ εg2 > > > > > G ¼ ðbd þ aεÞg > > < H ¼ g2 ad þ bε g > > > I ¼ d a2 þ b2 ag2 > > > > > 2 2 2 > > > J ¼ ε a þ b þ bg > > > : 2 d2 þ ε2 þ 2g2 ðbε adÞ þ g4 jdetj ¼ a2 þ b
(9)
where the parameters in Greek letters are expressed in terms of tire properties as
! 8 > EA 2 > > a ¼ mU n2 þ kt mU2 > 2 > > R0 > > > > > > b ¼ dt Un > > > ! > > < EA 2 g ¼ 2 2mU n > R0 > > > > ! > > > > EI 4 2EI T0 2 2 > > d ¼ 4n þ mU 2 n þ > > > R0 R0 R40 > > > : ε ¼ dr Un
(10) EA R20
þ
EI R40
T0 R20
! 2
mU þ kr
The study also presented equations to calculate extensional stiffness, bending stiffness, and bedding stiffness of the tread belt.
EA ¼ 4ðA Ac ÞGR 1 cot2 a cot4 a
(11)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
8 EA ðNtb Þ3 > > > < ðEIÞb ¼ t w wb 12 b b > 3 > > : ðEIÞ ¼ E w tt R t t 3 8 cos q0 þ q0 sin q0 > > > < kr ¼ sin q q cos q pi 0 0 0 > G t > s R > : kt ¼ þ p0 cot q0 ls
7
(12)
(13)
The damping coefficients for truck tires are approximately calculated using the following equations.
8 pffiffiffiffiffiffiffiffi > kr m > > < dr ¼ 10 pffiffiffiffiffiffiffi > > > : d t ¼ kt m 10
(14)
2.2.2.2. Air pumping. It was found that tread void compression occurs mainly through the lateral motion of the tread elements entering and leaving the contact patch, and not through the vertical compression of the tread rubber (Plotkin et al., 1980 [35]; Samuels, 1979 [36]). The reflection from the ground and tire sidewall plays an important role in noise generation by increasing the pressure by a factor of four, which is the geometric radiation factor G. Air pumping also depends on the flow blockage factor εb; e.g., for rib tires with continuous grooves, some of the displaced air moves within the grooves instead of being pumped out, in which case, εb z 0.35; however, for cross-bar tires, which generate more air pumping noise, εb z 1. To analyze the air pumping noise, the volume of a single void is calculated; this is given by
V ¼ wv d½lv lb hk0 þ ðlv þ lb Þhk
(15)
The equation for sound pressure for a monopole source is
p¼r
V€ 4pr
(16)
Assuming that k is a function of f, and that all the other parameters in V are constant, 2
d k V€ ¼ V 00 ðfÞU2 ¼ wv dðlv þ lb Þh 2 U2 df
(17)
The final equation for the sound pressure is
pðfÞ ¼
∞ Gεb r U 2 l X wv dhðlv þ lb Þ 2 n2 n2 1 ðCn sin nf þ Dn cos nfÞ 4pr R0 R0 n¼1
(18)
The sound pressure for a tread void with pitch ll, depth dl, and width wl is given by
pl ðtÞ ¼
ll dl wl p ðtÞ l0 d0 w0 0
(19)
where p0(t) is the sound pressure for a tread void with dimensions l0, d0, and w0. The sound pressure of air pumping for a full tread pattern is given by wt
Nv Z 2 ygi ðxÞ 1 X dx pðfÞ ¼ p0 f wt i¼1 R0
(20)
wt 2
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
8
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
2.2.2.3. Carcass vibration. The models are applicable only for steady uniform carcass on a smooth pavement. For the vibration of tire carcass on a textured surface, the circumferential coordinate was changed from f to s ¼ R0f, and ()0 denotes v/vs. For the contact patch edge excitation, the radial displacement of the belt is
w ¼ Weiðkt sutÞ ehs
(21)
8 u > > > kT ¼ > C > p > > > ! > > < 1 kr 2 2 Cp ¼ EIkT þ T0 þ 2 m kT > > > > > > > Cp kr > > > :h ¼ 4EIk2T þ 2T0
(22)
where
The motions at the leading edge fa and trailing edge fb due to the excitation of the tread pattern are
8 < € wðfa Þ ¼ P U2 f ð0ÞA0 ðfa UtÞ : wðf € b Þ ¼ P U2 f ð0ÞA0 ðfb UtÞ
(23)
For the pavement texture excitation, the load at point (x, y) is
zs ðx; yÞ ¼ Pð1 þ zðx; yÞÞ pr ðx; yÞ ¼ P 1 þ Dz
(24)
Then, the equation for the motion at the trailing edge fb due to the excitation of the pavement texture is
€ b Þ ¼ P U2 f ð0Þ A0 ðfb UtÞ þ Az0 ðs0 þ Rðfb UtÞÞR wðf
(25)
2.2.2.4. Sound radiation. The sound pressure from a moving surface S (Morse, 1970 [37]) is given by
Z pðrÞ ¼
! b 0 dS Gðrjr0 Þ V p$ n
(26)
S
Using Eberhardt's model (Eberhardt, 1979 [34]) and employing Green's function for a vibrating plate, considering the flat tread area of the tire as an infinite baffle, the final equation is obtained as wb
rW 2 iut u e pðr; uÞ ¼ 2p
Z 2 Z∞ wb 2
eika r ikt s hs e e ds r
(27)
0
2.2.3. Truck tire (Keltie, 1982) Keltie (1982) [38] modeled the truck tire as an infinitely long incomplete circular cylindrical shell, as shown in Fig. 2, to predict the sound power generated. The assumption of infinite length is based on the fact that the waves propagating circumferentially are effectively damped out before they can meet near the top of the tire. The tire is both a nonisotropic and nonhomogeneous material owing to the composite structure of the rubber and the laminated plies. However, in the model, it is considered as a homogeneous and isotropic material, considering only the ply region, as the ply contributes to most of the load-carrying ability. The boundary conditions of the tire model are pinned connections with the rim along the lines q ¼ 0 and q ¼ q0. Based on Flügge's thin shell theory, considering both bending and extensional effects, the equation of motion for the shell surface is given by Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
9
Fig. 2. Shell model geometry and coordinate system (source from Keltie, 1982 [38], Fig. 3; reprinted with permission from AIP Publishing LLC).
2
L11 4 L12 L13
L12 L22 L23
32 3
2 q ðs; q; tÞ 3 L13 uðs; q; tÞ 2 1 n2 x R 4 q ðs; q; tÞ 5 L23 54 vðs; q; tÞ 5 ¼ q Eh L33 wðs; q; tÞ qz ðs; q; tÞ
(28)
where L11, L12, …L33 are differential operators; u, v, and w are the deformations in the x, q, and z directions, respectively; qx, qq, and qz are the externally applied pressures in the x, q, and z directions, respectively; s is the normalized coordinate given by is the complex Young's modulus. To compute the acoustic radiation, the radial displacement s ¼ x/R; n is Poisson's ratio, and E needs to be obtained based on the radial load; the relevant equations are given below.
8 > q ðs; q; tÞ ¼ ejut qðsÞHðsÞ > > < z > > wðs; q; tÞ ¼ ejut wðs; qÞ ¼ ejut > :
∞ X
1 2 2 npq wn ðsÞ sin
n¼1
q0
(29)
q0
where q(s) and H(q) are polynomials that satisfy the differentiability and continuity conditions. The sound pressure p(x, r, q) is governed by the equation
! 2 v2 v2 1 v 2 v 2 þ r þ r þ þ K a pðx; r; qÞ ¼ 0 vr vr2 vq2 vx2
(30)
where r is the radial coordinate, and Ka is the acoustic wavenumber. The normal velocity of the shell surface vs(x, q) is related to the sound pressure p(x, r, q) through the boundary condition and Euler's equation, as given below.
jur0 vs ðx; qÞ ¼
vpðx; r; qÞ vr r¼R
(31)
where u is the angular frequency, and r0 is the air density. The surface acoustic intensity J(x, q) is calculated using the equation
Jðx; qÞ ¼
o 1 n * Re ½pðx; R; q; tÞ½vs ðx; q; tÞ 2
(32)
where Re indicates the real part of the complex value, * indicates the complex conjugate. The total radiated sound power P is given by
Z∞ Z2p
P¼
Jðx; qÞRdqdx
(33)
∞ 0
The tire parameters such as bending stiffness, inflation pressure, and structural damping were analyzed for their effects on the sound power level in 1/3 octave band and on the overall sound power level. 2.2.4. 2D ring (Kung et al., 1986) Kung et al. (1986) [39], and Huang and Soedel (1987) [40,41] considered the tire belt as a two-dimensional rolling ring, and the sidewalls as an elastic foundation with radial and tangential stiffness. The equations of motion are given below (Huang and Soedel, 1987 [40]). Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
10
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
8 p K 0 00 D 000 p0 2 > 0 00 00 0 > €q þ 2Uu_ r ¼ qq ðq; tÞ þ u 2u þ k u þ u r h U þ u þ r h u u u þ > q q r r r q q q < a4 a a a2 >
D K p0 p0 > 2 000 0 0 00 > € _ : 4 u000 u þ k u þ u r h U þ u þ r h u 2 U u þ u þ þ 2u r r r r q ¼ qr ðq; tÞ r r q q q a a a a2
(34)
where kr is the radial stiffness, kq is the tangential stiffness, ur is the radial displacement, uq is the tangential displacement, qr and qq are the external forces per unit area in the radial and tangential directions, respectively, D [¼ Eh3/12(1-n2)] is the bending stiffness, K [¼ Eh/12(1-n2)] is the membrane stiffness, E is Young's modulus, r is the ring density, n is Poisson's ratio, a is the mean radius of the tire, h is the thickness of the tread, p0 is the internal pressure, and U is the rotational speed. The dots and primes denote differentiation with respect to time t and circumferential coordinate q, respectively. The modal analysis showed good agreement between theory and measurement in the frequency range of 0e400 Hz. At higher frequencies, when the wavelength is close to the tread width of the tire, the model fails, because the tire tread band should be modeled as a 3D plate instead of a 2D ring. The naming convention for the mode shapes can be found in the work by Wheeler et al. (2005) [42]. €hm (1966) [43], where the tire was modeled as a ring under The fundamental 2D circular ring model was developed by Bo the tension caused by inflation pressure. 2.2.5. Various mechanisms (Heckl, 1986) Heckl (1986) [44] investigated multiple tire noise generation mechanisms, including tire vibrations, air pumping, pavement texture impact, stick/slip, and sound radiation from the tire. The circular ring model for tire vibrations, excluding tread block motion, is based on the equation proposed by Boehm (1966) [43].
8 ES B 000 00 0 0 > > < 2 ðu þ v Þ 4 ðv þ v Þ ¼ rSu€ þ kT u a a > > : T0 ðv00 þ vÞ ES ðu0 þ vÞ B ðv0000 þ v00 Þ ¼ rSu€ þ ka v q a2 a2 a4
(35)
where a is the radius, E is Young's modulus of the tire material, B is the bending stiffness, r is the tire density, kT is the stiffness of tangential bedding, ka is the stiffness of radial bedding, T0 is the tension, S is the cross-sectional area, q is the external driving pressure in the radial direction, v is the radial displacement, and u is the tangential displacement. The time derivatives are indicated by dots and derivatives with respect to the angle f by primes. The circular ring model appears to be more useful than the commonly used modal analysis, especially when the frequencies and/or the damping levels are high. However, the wavelength on the tire surface must be longer than approximately four times the tread thickness h or the tread block distance d (whichever is larger). For typical tires (a z 0.28 m, d z 25 mm, cI z 80 m/s), this gives an upper frequency limit of 800 Hz. Above this frequency, the tread blocks have to be considered individually. The air pumping model is based on the relationship proposed by Hayden (1971) [29].
P¼
2
r0 vq 4pc0 vt
(36)
where P is the sound power, r0 is the air density, c0 is the speed of sound in air, q is the volume flow from the grooves, and t is the time. As q f V, where V is the vehicle speed, and v/vt f V, the relationship between the air pumping sound power and vehicle speed is obtained as P f V4. The pavement texture impact model represents the tread blocks by small springs of stiffness s. Each of them, indexed as n, generates a force given by
Fn ¼ sðvT n vRn Þ
(37)
at the point fv which is determined by the difference between the tire displacement vT and the road surface amplitude vR at the vth point. Then, the resulting deflection at point fi is
vTi ¼
N X
Fn gðfi fn Þ
(38)
n¼1
where N is the number of springs, and g is the Green's function for a unit force. For the direct air-borne sound excitation of a blank tire on a rough road, another theory was proposed (Ronneberger, 1984 [45]).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
11
Stick/slip at the trailing edge was discussed, but no detailed quantitative theory was provided. The author also indicated that spatial changes in the bending stiffness of the tire from one tread block to another are transformed into temporal changes, and therefore they can also be found in the sound signal, but for a slick tire (homogeneous), no such effect is observed. 2.2.6. Tire vibration (Bremner et al., 1997) Bremner et al. (1997) [46] developed a tire vibration model to correlate the vibration-radiated noise with tire construction and material properties. The tire was modeled as an orthotropically stiffened and pressurized cylindrical shell with boundary condition as simply supported. This model was based on the model developed by Mikulas and McElman (1965) [47] and extended by Rennison et al. (1979) [48] (geometric tension stiffness included). This model included membrane waves in the lower frequency range [49] and bending waves in the higher frequency range. Statistical energy analysis (SEA) with modal joint acceptance formulations was used to calculate the wave speed and radiation efficiency. All the main tire design parameters were analyzed for sensitivity study. Wavenumber decomposition was used to determine the dispersion characteristics of a stationary tire under radial point force excitation for experimental validation. The mode shapes for the tread and sidewall, considered as a uniform construction, are as follows.
8 Jmn ðx; qÞ ¼ sinðkm xÞsinðkn RqÞ > > > > < mp km ¼ L > > > n > :k ¼ n R
(39)
where L is the circumference of the wheel rim, R is the smaller radius of the tire cross section, and m and n are mode indexes. It was found that below 300 Hz, tension stiffness has a dominating effect, where the natural frequency is proportional to the circumferential wavenumber. Above 300 Hz, bending stiffness begins to show a dominating trend, where the natural frequency is proportional to the square of the circumferential wavenumber. SEA gives the following relationship. in
PDu ¼ uðht þ hrad ÞEt;Du
(40)
where Pin Du, with a bar above, is the time-averaged power input to the tire in each frequency band Du, u is the temporal frequency of vibration, ht is the structural and material damping coefficient, and hrad is the radiation coupling loss factor (much smaller than ht). Without considering the direct vibration field (valid up to 700 Hz), the energy level given by SEA is
Et;Du ¼ Mt 〈v2t 〉sp;Du
(41)
where Mt is the tire mass;
sp,Du, with a bar above, is the space-averaged and time-averaged reverberant vibration level (particle velocity in the sound field). The power input is applied based on the study by Remington et al. (1983) [50], in which the contact point velocity is determined by the roughness profile of the two contact surfaces and their relative contact point impedances. Assuming that the impedance of the pavement is much larger than that of the tire, the power input is given by
in
PDu ¼ v2i;Du Re Zii∞
(42)
where v2i,Du, with a bar above, is the mean squared vibration velocity at the tire contact point, and Z∞ ii is the point “infinite impedance” of the tire. The term v2i,Du, with a bar above, is given by
v2i;Du ¼
u2 〈F2road ðkÞ þ F2tread ðkÞ〉 V
$Du
(43)
where F2road(k) and F2tread(k) are the power spectral density (PSD) values of the road roughness and tread roughness (the two spectra are uncorrelated), respectively; k is the spatial roughness wavenumber defined by
k¼
u V
(44)
where V is the tire rolling speed. The model showed good agreement with the experimental results for far-field noise measurements (sound pressure level or SPL at 7.5 m), including the effects of rolling speed, rubber damping, inflation pressure, tread thickness, and tread materials.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
12
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
2.2.7. Porous pavement (Berengier et al., 1997) rengier et al. (1997) [51] studied the acoustical performance of a porous pavement, and developed analytical models to Be describe the surface impedance and sound propagation. Two different models were developed for the impedance characterization, i.e., a phenomenological model and a microstructural model; both models included viscous and thermal effects in the porous structure. The two models have a general formulation as given below.
8 1
> rg ðuÞ 2 > > u k ¼ > > < Kg ðuÞ h i1 > 2 > > rg ðuÞKg ðuÞ > > :Z ¼ c
(45)
U
where k is the complex wavenumber, Zc is the characteristic impedance, u is the angular frequency, U is the porosity, rg is the complex dynamic density, and Kg is the bulk modulus. In the phenomenological model, three physical input parameters are included: airflow resistivity of the porous structure Rs, porosity of the air-filled connected pores U, and tortuosity q2. The complex dynamic density and bulk modulus are given by
8 fm 2 > > 1 þ i r ð u Þ ¼ r q > g 0 > f > < 31 2 > > ð g 1Þ > 4 > 5 > : Kg ðuÞ ¼ gP0 1 þ 1 i ff
(46)
q
where r0 is the air density, g is the specific heat ratio, P0 is the atmospheric pressure, i is the imaginary unit, f is the frequency; fm and fq describe the viscous and thermal dependencies, and are given by
8 > > > < fm ¼ > > > : fq ¼
URs 2pr0 q2
(47)
Rs 2pr0 Npr
where Npr is Prandtl number. In the microstructural model, five physical input parameters are included: besides Rs, U, and q2, the other two pore shape parameters are sr, associated with viscous dependency, and sk, associated with thermal dependency. The complex dynamic density and bulk modulus are given by
8 Rs U F l p > 2 > > < rg ðuÞ ¼ r0 q þ i
u
> ðg 1ÞTðLk Þ 1 > > : Kg ðuÞ ¼ gP0 1 þ 2 Lk
(48)
where F(lp) is a function of sr, and T(Lk) is a function of sk. It was found that the phenomenological model with only three input parameters is accurate enough for acoustic characterization. In addition, it was found that Rs and U are more important than q2. The models can also be used for the detection of pavement clogging. 2.2.8. Wavenumber decomposition (Bolton et al., 1998) Bolton et al. (1998) [52] represented the tire vibration in the form of propagating waves instead of the traditional modal form. The authors claimed that the wavenumber domain form might be more closely related to the physics of wave propagation in tires. The radial vibration of the tire carcass is given by
ur ¼
N X
Am Jm ðsc ; uÞe±ikqm s
(49)
m¼1
where ur is the radial velocity, m is the wave type number, N is the number of contributing wave types, Am is the wave type participation coefficient, Jm(sc, u) is the cross-sectional mode shape, s is the path variable in the circumferential direction, sc is the path variable in the cross-sectional direction, and kqm is the circumferential wave number (real part indicating the propagation speed, and imaginary part indicating the spatial damping). The wave type number m ¼ 1 corresponds to the fast
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
13
extensional mode (180 m/s), while m ¼ 3, 5, and 7 correspond to slow flexural modes (60e80 m/s), as shown in Fig. 3. In this figure, each curving trajectory indicates a single wave type (slope being the propagating speed). The model on the radial tire velocity can then be applied as the boundary condition to the acoustic radiation model. The study investigated a stationary passenger car tire suspended horizontally and driven radially by a shaker at a point on the tread band. It was reported that at low frequencies, a small number of slowly propagating flexural modes dominate, while at higher frequencies the fast modes of the tread band extension are dominant. The type of gas (helium, carbon dioxide, and air) inflated into the tire cavity was seen to have no effect on the carcass vibration, while the inflation pressure (20 psi and 40 psi) had major effects. The cut-on frequencies increased with inflation pressure. It was also found that only the modes m ¼ 3 and n ¼ 1 (cut-on circumferential mode of the first propagational flexural mode) can deliver a net transverse force to the wheel hub and contribute to the transmission of structure-borne vibration. 2.2.9. Orthotropic plate (Kropp, 1999) The orthotropic plate model was developed by Dr. Kropp (1999, Department of Applied Acoustics, Charlmers University of Technology, Sweden) [53]. This model analyzes a smooth tire rolling on a rough pavement. It also consists of three submodels very similar to the one in the study by Kropp (3D two-plate that will be discussed later) (1992) [25,26], and addresses only the structure-borne noise. The differences are the following: (1) In the first submodel (tire model), the tire was simplified as a 2D orthotropic plate for vibration calculations. (2) In the second submodel (contact model), a bedding model was used for the tread, together with a time iterative solution method, to solve the nonlinear problem. (3) In the third submodel (sound radiation model), not only was the vibration of the tire belt considered, but the geometrical condition of the horn effect and local deformations of the tread were also included by using the multipole synthesis method (Cremer, 1984 [54]; Heckl, 1989 [55]). The advantage of this model is that it maintains sufficient accuracy as in the case of the Kropp (3D two-plate) model, and reduces computational effort. As a result, the measurement and calculation showed better agreement over a broader frequency range (125e2500 Hz). A brief introduction to the three submodels is given below. The tire model considers the belt and the two sidewalls as flat plates with different bending stiffnesses in the longitudinal and lateral directions, as illustrated in Fig. 4. The differential equation for the vertical motion of the plate is given by
! 8 00 00 d2 x d2 x d4 x d4 x d4 x > > þ 2 þ Bx 4 þ 2Bxy 2 2 þ By 4 m u2 x þ sx ¼ F0 < T0 2 dx dy dx dx dy dy > qffiffiffiffiffiffiffiffiffiffi > : Bxy ¼ Bx By
(50)
where T0 is the tension caused by the inflation pressure; Bx, By, and Bxy are the circumferential, axial, and cross stiffnesses, respectively; x is the vertical displacement, s is the spring stiffness of the elastic foundation, m00 is the mass per unit area, and F000 is the force acting per unit area. The submodel was validated based on measurements such as the complex point mobility (or admittance, i.e., velocity divided by force) on a smooth, unloaded, stationary tire. A drawback of this model is that it neglects the tire curvature and thickness of the plate, and hence there is a shift in the resonance frequencies below the ring frequency of 400 Hz (circular structure), and large errors at frequencies above 3000 Hz (wavelength is comparable to plate thickness) (Übler, 1994 [56]).
Fig. 3. Frequency-wavenumber characteristics of tire carcass vibration (source from Bolton et al., 1998 [52], Fig. 8; reprinted with permission from Dr. J. Stuart Bolton of Purdue University, USA).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
14
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Fig. 4. Schematic of an orthotropic plate tire model (source from Kropp, 1999 [53], Fig. 2; reprinted with permission from Dr. Jeanette Brooks on behalf of Inderscience Enterprises Limited).
The contact model uses Green's function to calculate the displacement response at a given location as the result of a force applied at another arbitrary location. The discrete time domain equation is
xe ðNDtÞ ¼
X
Fm ðN DtÞgm;e ð0Þ þ
m
X N1 X m
Fm ðNDtÞgm;e ½ðN nÞDt
(51)
n¼0
where xe(t) is the displacement at the discrete point e at time t, Fm(t) is the driving force at the discrete point m at time t, gm,e(t) is the Green's function representing the relationship between the input at the discrete point m and the output at the discrete point; Dt is the time increment, and N is the time step index. The local deformation of the tread is not considered in the Green's function, but will be compensated for later. The sound radiation model assumes two multipoles located symmetrically on either side of the road surface, as illustrated in Fig. 5. The sound pressure at point E is
pðxE ; yE ; tÞ ¼
X n
ð2Þ
an Hn ðkr1 Þejnf1 þ
X
ð2Þ
bn Hn ðkr2 Þejnf2
(52)
n
where r1, f1, and r2, f2 are the coordinates of point E(xE, yE) with respect to the two coordinate systems. H(2) n is the Hankel function of the second kind and nth order; an and bn are unknown coefficients; k is the wavenumber, j is the imaginary unit, and n is the circumferential mode index. The particle velocity at point E is given by
1 X ð2Þ0 an kHn ðkr1 Þejnf1 jur n 8 9 > sin f1 r1 b cos f1 > > > = < r sin f 1 1 r2 1 X r b cos f1 ð2Þ0 r23 ð2Þ0 rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ rn a n k 1 Hn ðkr2 Þ þ jnHn ðkr2 Þ ejnf2 2 > > jur n r2 > > r1 sin f1 ; : 1
vE ðr1 ; f1 Þ ¼
(53)
r2
0 where H(2) n (kr) is the derivative of the Hankel function with respect to its argument, b is the distance between the multipoles, rn is the reflection factor of the road surface for normal incidence, r is the air density, and u is the angular frequency. The
Fig. 5. Schematic of the two multipoles for the sound radiation model (source from Kropp, 1999 [53], Fig. 6; reprinted with permission from Dr. Jeanette Brooks on behalf of Inderscience Enterprises Limited).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
15
boundary conditions, i.e., the velocities normal to the tire surface, vzn, are determined by the velocity distribution calculated using the tire model and contact model. To calculate the sound power P(u), the tire was modeled as a finite cylinder of width b, as shown in Fig. 6. The equation for P(u) is
PðuÞ ¼
Z∞ ∞ rca X jvrn ðu; kz Þj2 sn ðu; kz Þdkz 8p2 n¼∞
(54)
∞
where a is the tire radius, c is the speed of sound in air, kz is the wavenumber in the axial direction, and vrn(u, kz) is the Fourier coefficient of the radial velocity. The equation for vrn(u, kz) is given by vrn(u, kz) ¼ vrn(u)$sin(kzb)/b. sn(u, kz) is the sound radiation efficiency, given by
sn ðu; kz Þ ¼
2u 1 1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 apc u2 2 ð2Þ0 2 kz H c2 a uc2 k2z n
(55)
The full model was validated using the laboratory drum method (DR), as illustrated in Fig. 7, by measuring three parameters: (1) The radial accelerations were measured using accelerometers mounted in the tread grooves (tire vibration measurement, TVM) in the Lagrange reference; (2) The local velocities were measured using a laser Doppler vibrometer (LDV) in the Euler reference; (3) The radiated sound pressure level was measured using a microphone. The first
Fig. 6. Finite cylinder model (source from Kropp, 1999 [53], Fig. 8; reprinted with permission from Dr. Jeanette Brooks on behalf of Inderscience Enterprises Limited).
Fig. 7. Schematic of measurement facilities (source from Kropp, 1999 [53], Fig. 11; reprinted with permission from Dr. Jeanette Brooks on behalf of Inderscience Enterprises Limited).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
16
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
measurement showed higher values than those obtained by calculation, but the other two measurements showed good agreement. However, deviations were found to occur above 1500 Hz, because the local deformation of the tread, which acts like a small volume source, becomes more dominant than the belt vibrations. After adding the radiated sound, which is calculated as the integral over the displaced volume, the total third octave band showed very good agreement in the range of 125e2500 Hz. In the present authors' opinion, the Kropp model (orthotropic plate) might be currently the most advanced model, with a suitable approach for tire vibration modeling in tire noise studies. 2.2.10. Shell (Kim and Bolton, 2002) Kim and Bolton (2002) [57,58] analyzed the potential of a rotating tire to radiate sound from static tire dispersion using kinematic compensation. The tread band of the tire was modeled as a simply supported, rotating circular cylindrical shell (Matsubara et al., 2015 [59]) having a finite width to identify the effects of rotation. The governing equations for the tread band model are
D v v ¼ þU Dt vt vf
(56)
8
Dux D2 ux > > > þ rh ¼ qx ðx; f; rÞ Lx ux ; uf ; ur þ l > > Dt > Dt 2 > > > 2 <
Duf D uf Dur þ rh U2 uf ¼ qf ðx; f; rÞ Lf ux ; uf ; ur þ l þ 2U 2 > Dt Dt Dt > > > > 2 >
> Du D ur Duf > 2 > : Lr ux ; uf ; ur þ l r þ rh 2 U U u r ¼ qr ðx; f; rÞ Dt Dt Dt 2
(57)
8
vNxx 1 vNfx > > > > Lx ux ; uf ; ur ¼ vx a vf > > > > > <
vNxf vNff Qfr Lf ux ; uf ; ur ¼ vx vf a > > > > > r 2 2 > > > L u ; u ; u ¼ vQxr 1 vQfr þ Nff N r v ur Nff v ur > : r x f r xx 2 2 a vf vx a vx a vf2
(58)
where the subscripts x, f, and r denote axial, circumferential, and radial directions, respectively, of the tire, and the superscript r denotes the residual force. U is the angular rotational speed, f is the circumferential angle in the reference frame, and u is the displacement in the corresponding direction; N and Q are the resultant normal and shear forces, respectively; q is the external force applied in the corresponding direction, r is the density of the tread band, h is its thickness, l is the damping constant, and a is the tire radius. Nrff is the resultant circumferential normal force given by r Nff ¼ ap þ rha2 U2
(59)
where p is the inflation pressure. The linear operators Li(i ¼ x, f, r) are related to the system's stiffness. The rotation of the tire was found to have two effects: stiffening of the tread band and “tilting” of the dispersion curves. However, at normal speeds, inflation pressure has greater influence on the stiffness than rotation. In contrast, the second effect (kinematic tilting effect) was found to be significant. The kinematic relationship is given by
~ mn ¼ umn nU u
(60)
where ῶmn is the natural frequency in local (Lagrangian, tire-fixed) coordinates, umn is the natural frequency in global (Eulerian, vehicle-fixed) coordinates, and n is the wave mode number in the circumferential direction. 2.2.11. Coupled cavity (Molisani et al., 2003) Molisani et al. (2003) [60] modeled the tire as two shells of revolution and two annular plates, where only the outside shell is flexible, i.e., the tire sidewalls and rim are assumed to be rigid. The boundary conditions for the outer shell were assumed to be simply supported. This model provided an efficient tool to investigate the physical coupling mechanisms between the acoustic cavity and the tire structure without complicated computations such as finite element analysis (Richards, 1991 [61]), where these mechanisms are often hidden in the large numerical models. For the structural equation, the displacement vector of the outer shell was obtained by solving the eigenvalue problem of the self-adjoint Donnel-Mushtari operator. For the acoustic equation, a finite cylindrical cavity with rigid boundary conditions
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
17
was developed. This model investigated the interaction between the acoustic mode and the structural radial mode. The equation of motion for the system, considering fluid loading, is
8 9 8 9 8 9 0 = nh io< u = < 0 = < 2 0 Lc U ½I v ¼ 0
: ; : ; : 0 ; p a; x0 q w fw
(61)
where Lc is the shell differential operator, U is the normalized frequency, I is the identity matrix, {u, v, w}T is the displacement vector, where the components u, v, and w are the displacements in the axial, tangential, and radial directions, respectively; fw is the vertical load, p(a, x0 , q0 ) is the cavity acoustic pressure acting on the tire. The equation of the coupled system is
n o ½K U2 ½M fAmn g ¼ ffmn g U2 ½afAmn g
(62)
where [K] and [M] are diagonal matrices, and their elements represent the modal stiffness and the mass of the radial modes, respectively; {Amn} is the vector of modal amplitudes, {fmn} is the vector of modal forces, and [a] is the fluid coupling matrix. [a] indicates the extent to which the acoustic mode will affect the structural mode. It was shown that the tire cavity resonance could largely increase the forces at the spindle, and thus increase the vehicle interior noise. It was also found that relatively high damping of the tire structure does not effectively damp out the energy in the acoustic resonance. However, the tire was assumed to be stationary without rotation, and the external excitation was assumed to be a point force acting normal to the shell, so it might not be able to represent the actual cavity resonances when the tire is in motion on a road. 2.2.12. Waveguide (Kim et al., 2007) Kim et al. (2007) [62] developed an analytical model for the tire vibration and structure-borne noise. Three models were developed to represent different vibrations or wave types (waveguide characteristics) at different frequencies: (1) springmass model for low frequencies, below 80 Hz; (2) elastically supported beam model for frequencies in the range of 80e300 Hz; (3) cylindrical shell for frequencies in the range of 300e2000 Hz. Below 80 Hz, the spring-mass model considers the tire sidewall as a spring, and the tread band as a mass. Four modes are found to be mostly related to sidewall stiffness. The eigen-frequencies of the first radial, lateral, and tractive modes, ur, ux, and ut, respectively, are given by
8 rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > pR0 ðKr þ Kt Þ > > u ¼ > r > > M > > > rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi < 2pR0 Kx ux ¼ > M > > > rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > > > > > u ¼ 2pR0 Kt : t I
(63)
where R0 is the tire radius, Kr, Kx, and Kt are the sidewall spring constants in the radial, lateral, and tractive directions, respectively; M is the tire tread band mass, and I is the polar moment of inertia of the tire tread band. For the frequency range of 80e300 Hz, the tire tread band is modeled as a tensioned beam (R0 ¼ ∞) supported elastically by the sidewall springs with flexural waves propagating in the circumferential direction. The dispersion relationship of the free flexural waves is given by
8 2 > U ¼ b2 ðR0 kÞ4 þ g2 ðR0 kÞ2 þ U2r > > > > >
u
0 > > > > h > > : b ¼ pffiffiffiffiffiffi 12R0
(64)
where U is the non-dimensional frequency, b is the non-dimensional thickness, h is the tread band thickness, and u0 is the ring frequency. For frequencies over 300 Hz, the second and third radial modes occur, and the one-dimensional waveguide turns into two-dimensional waveguide, i.e., cylindrical shell/plate. Waveguide behavior and shell curvature are the two important properties that determine the characteristics of the 2D flexural waves. The approximate dispersion relationship is given by
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
18
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
8 2 n2 ð4 nÞ 2 n m4 2 2 > 2 2 2 > m U ¼ 1 n þ b þ n > 2 > 2ð1 nÞ > < m2 þ n2 > > > m ¼ kz R 0 > > : n ¼ ks R0
(65)
where n is Poisson's ratio; and kz and ks are circumferential and axial wavenumbers, respectively. The cross-sectional resonant frequencies (cut-off frequencies) are obtained by applying kz ¼ 0 in the above equation. The dispersion relationship accounting for the shell curvature that results in the coupling between radial, axial, and circumferential motions is given by the following equations.
8 > U2 ¼ b2 ðR0 kcs Þ4 þ cos4 f > > > > < 2 kcs ¼ k2z þ k2s > > > > 1 ks > : f ¼ tan kz
(66)
It was also concluded that the tire waves discussed above are inefficient sound radiators, since their wavenumbers are larger than the acoustic wave numbers. Acoustically excited wave motion of the tire tread band is an important factor for sound radiation. 2.2.13. Cavity (Feng et al., 2009) Feng et al. (2009) [63] investigated the effects of tire rotating speed and tire load on the cavity resonance mechanism. The lowest two cavity modes (splitting into two modes because of the tire load) for a non-rotating tire are defined as horizontal (fore/aft) acoustic mode and vertical acoustic mode. The equations of motion for the two modes are
8 < X€ þ 2zu0 X_ þ u2 þ af0 X ¼ 0 0 : Z€ þ 2zu Z_ þ u2 þ bf Z ¼ 0 0 0 0
(67)
where X and Z are the coordinates of the center of mass of the air molecules, u0 is the natural frequency of the acoustic cavity of the unloaded tire, z is the damping ratio, f0 is the static load on the tire outer surface in the Z direction; a (negative) and b (positive) represent the effects of tire load on the natural frequencies of the two acoustic modes. The equations for the natural frequencies of the two modes are
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 8 < u ¼ u2 þ af H 0 0 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : 2 uV ¼ u0 þ bf0
(68)
where uH and uV (uH < uV) are the resonance frequencies for the horizontal and vertical directions, respectively. For a rotating tire with a rotation angle f, the transformation between the tire-fixed coordinate (X, Z) and the vehicle-fixed coordinate (x, z) is given by
x ¼ X cos f þ Z sin f z ¼ X sin f þ Z cos f
(69)
It may be noted that the rotation rate is slow compared to the lowest acoustic mode frequencies. The equations for the natural frequencies for the rotating tire are
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 8 2 h i > > aþb ab 2 2 2 2 > > f f u ¼ u þ U þ þ 2ða þ bÞf þ 4 u U2 > 0 0 0 1 0 0 < 2 2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > 2 h > i > a þ b ab > > u2 ¼ u2 þ U2 þ : f0 þ f0 þ 2ða þ bÞf0 þ 4u20 U2 2 0 2 2
(70)
where U is the rotation speed of the tire. Even though the natural frequencies of the two modes vary with U, they have the constant average frequency ῶ0 as given below, where the peak SPL occurs.
~0 ¼ u
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 1 2 u20 þ ða þ bÞf0 ¼ u þ u2V 2 2 H
(71)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
19
Further, Lagrange's equations were implemented into the model (Feng and Gu, 2011 [64]). 2.2.14. 3D multilayer (O'Boy and Dowling, 2009) O'Boy and Dowling (2009) [65] provided a multilayer viscoelastic cylindrical representation of the tire belt, consisting of thin orthotropic layers of different materials of varying thicknesses laid on each other, as illustrated in Fig. 8. It overcame the disadvantage of inaccuracy at high frequencies when the shortest wavelength in the tire becomes comparable to the thickness. Each layer is considered to be isotropic in the circumferential and axial directions, but orthotropic in the radial direction. The model assumed the tire side walls to be pinned, and tire belt to be periodic in the axial direction. The model made it possible to understand the vibration characteristics of a stationary tire, including resonant frequencies, mode shapes and wave speeds, using the parameters of the belt layer structure (density, Young's modulus, Poisson's ratio, and thickness), defined solely by design data without experimental testing. The stress s and displacement u on the lower surface of layer j are related to the coefficients within that layer (O'Boy and Dowling, 2009 [65]).
s~ G ¼ W S C ~ u H r rj rj rj r j
(72)
j
where ~ denotes Fourier transform in the frequency domain, G and H contain all the information on how the material properties in the viscoelastic layer affect the stresses and displacements for all wavenumbers and frequencies, W has only diagonal terms depending on the Wronskian expressions for the Bessel functions, and S is a summation matrix. Furthermore, the components of stress and displacement on one face of the elastic layer can be related to the components on the other face, as given below.
s~ ~ u
¼
rNþ1
N Y
Xj
s~ ~ u
(73)
r1
j¼1
where Xj is the overall material matrix of layer j, which is a function of frequency, wavenumber, material properties of the layer, and its inner and outer radii. Xj is expressed as
Xj ¼
G H
rjþ1
Wrjþ1 SMrjþ1;j S1 W1 rj
G H
1 (74) rj
where Mr is the scaling factor that depends on the thickness of the layer. The most useful vibration data for noise prediction, ~ as given below. the displacement u, can be obtained from the inverse Fourier transform of u
uðr; t; q; yÞ ¼
Z∞ Z∞
1 3
ð2pÞ
∞ ∞
∞ X
~ r; n; ky ; u einq eiky y eiut dky du u
(75)
n¼∞
where r is the radius of the belt layer; t is the time, q is the circumferential angle, and y is the axial position; u, n, and ky are the corresponding variables in the frequency domain, i.e., angular frequency, angular order, and axial wavenumber, respectively. The model can provide the displacement or velocity response of the tire belt when excited by a radial or tangential force. The response can be displayed in both the frequency-wavenumber and time-spatial domains. It was shown that the dominant flexural wave under radial excitation travelled at 50e65 m/s, while the compression wave under tangential excitation travelled at 150 m/s.
Fig. 8. Complete viscoelastic cylindrical model of the tire belt (source from O'Boy and Dowling, 2009 [66], Fig. 1; reprinted with permission from Elsevier).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
20
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
O'Boy and Dowling (2009) [66] also introduced an approximation using a flat bending plate for the tire belt. The vertical displacement w(x, y, t) at position (x, y) at time t is governed by (Graff, 1991 [67]) the following equation.
pffiffiffiffiffiffiffiffi v v2 w TV2 w DV4 w þ p þ hp DM V2 w ¼ M 2 vt vt
(76)
where T is the in-plane tension with the unit N/m, p(x, y, t) is the load per unit area, M is mass per unit area (area density), hp is a constant representing additional damping, and V is the gradient operator resulting in a vector like v/vx1 and v/vx2; V2 is the Laplace operator resulting in a scalar like v2/vx21 and v/vx22; D is the bending stiffness given by
D¼
Eh3
12 1 n2
(77)
where E is Young's modulus, and n is Poisson's ratio. This corresponds to the case for a tire belt without the tread rubber layer. For the case with tread layer included, the equation becomes
r v2 TV w þ V 2 k m vt 2 2
2
!
rh3 v2 12 vt 2
! DV
2
pffiffiffiffiffiffiffiffi v w þ ks w hp DM V2 w þ vt
! DV2 r h 2 v2 v2 w 1 2 þ p¼M 2 vt k mh 12k2 m vt 2
(78)
constant that accounts for where k is the Timoshenko shear coefficient that accounts for shear deformation, m is the Lame rotary inertia, ks is the stiffness of the equivalent distributed springs representing the tire sidewalls, and r is the rubber density. Then, the transfer function in the frequency domain between displacement w (with tilde above) and pressure p (with tilde above) is expressed as
I0 k2y þ I1 ~ w ¼ ~ ðu;n;k Þ J1 k4y þ ðJ2 þ I2 Þk2y þ ðJ3 þ I3 Þ p y
(79)
where J is the material term associated with the bending plate, I is the correction term that accounts for the shear deformation and rotary inertia. These are expressed by the following equations.
8 > > > > J1 ¼ D > > > > > pffiffiffiffiffiffiffiffi > 2Dn2 > > J2 ¼ 2 þ T þ iuhp DM > > > R > > > pffiffiffiffiffiffiffiffi > 2 > > Dn4 Tn2 iuhp DMn > > J ¼ þ þ þ ks Mu2 > 3 > 4 2 2 > R R R > > > > > D > > < I0 ¼ 2 k mh > > > > Dn2 u2 rh2 > > I1 ¼ 1 þ 2 > > > k mhR2 12k2 m > > > > 3 > > > Dr > 2 rh > þ I ¼ u > 2 > 12 k2 m > > > > > > 3 > Dr n2 u4 r2 h3 > 2 rh > > þ ¼ u þ I > 3 > 12 k2 m R2 > 12k2 m :
(80)
A numerical method was also presented to determine the parameters for the equivalent plate model which closely matches the multilayer viscoelastic cylindrical model. 2.2.15. Green's function (Arteaga, 2011) Arteaga (2011) [68] derived a closed-form expression for the Green's function for a loaded rotating tire expressed in the Eulerian reference coordinate, based on the modal-arbitrary Lagrangian-Eulerian (M-ALE) approach (Lopez et al., 2007 [69]). It addresses the contact problem directly in the Eulerian reference frame, and includes local tire softening due to nonlinear deformation without increasing computational cost. The expression for the Green's function gjl between the excitation at a given location and direction l, and the response at another location and direction j in the Eulerian reference frame is given below. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
gjl ðU; tÞ ¼
N X
"
Fjk
2N X
PN vkr
p¼1 wpr Flp sr t
r¼1
k¼1
ar
21
#
e
(81)
where U is the rotating velocity of the tire, t is the time, Fjk and Flp are the corresponding elements of the eigenvector matrix, vkr and wpr are the corresponding elements of the relative eigenvectors of transformed uncoupled first order differential equations, sr is the corresponding eigenvalue (same for vkr and wpr), and ar is an intermediate variable. It may be noted that the only details needed to compute these Green's functions are the eigenvalues and eigenvectors of the statically loaded tire in the tire reference frame and the mass matrix, which can be easily obtained from standard finite element modeling. The full excitation response of the tire is then calculated from the convolutions between the Green's functions and the contact forces. It was shown that the modal density increases as the rotation speed increases for both the unloaded and loaded tires. 2.2.16. Coupled Mode (Bolton and Cao, 2013) Bolton and Cao (2013) [70] investigated the coupled tire structural-acoustic mode. The tire tread band is unwrapped and modeled as a tensioned 1D string, while the rim is modeled as a hard wall, as illustrated in Fig. 9. In the figure, R is the tire radius, H is the tire section height, and L is the circumferential length. The tire cavity lies between the string and the hard wall. The air that rotates with the tire in the cavity is assumed to be uniform with inviscid flow, governed by the following linearized acoustic wave equation.
v2 p v2 p 2 v2 p v2 p ¼ c v2 þ c2 2 2v vxvt vt 2 vx2 vy
(82)
where p is the sound pressure, c is the adiabatic speed of sound, v is the flow speed, and t is the time; x and y are the circumferential and radial coordinates shown in Fig. 9. The wave equation for a tensioned moving string is given below.
rL
v2 x v2 x þ 2v 2 vtvx vt
!
v2 x þ rL v2 T ¼0 vx2
(83)
where rL is the linear density of the string, x is the radial displacement of the string in the y-direction, and T is the string tension force; v is the speed of the moving string, assumed to be the same as the air flow speed mentioned above. The solutions assumed are given below.
(
p ¼ Aejky y þ Bejky y ejkx x ejut
(84)
x ¼ Cejks x ejut
where A, B, and C are the amplitudes, j is the imaginary unit, kx and ky are the acoustic wavenumbers of the air flow in the xand y-directions, respectively; ks is the wavenumber of the string, and u is the angular frequency. Assuming kx ¼ ks, and with boundary conditions applied to the fluid-rigid body coupled system, the characteristic equation is
u2 2vukx þ v2
T
rL
k2x ¼
r0 u2
rL ky tan ky H
(85)
The dispersion equation is
Fig. 9. Tire-rim assembly modeled as tensioned string-hard wall (source from Bolton and Cao, 2013 [70]; reprinted with permission from Dr. J. Stuart Bolton of Purdue University, USA).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
22
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
u2 c2
¼
1
v2 2 vu k þ k2y 2 2 kx c2 x c
(86)
In this study, frequency-split phenomenon was observed during tire rotation in the coupled mode. The phase velocity of each mode approximately equals the associated static phase velocity plus/minus flow velocity. It was found that the plane wave mode is significantly influenced by the coupling effect, while the higher modes are less influenced. In the further study, Cao and Bolton (2015) [71] improved this model significantly by considering the tread band as a 2D ring, and including flexural and longitudinal waves. The displacement of the ring occurs in both circumferential and radial directions. The solution of acoustic pressure in the air cavity is assumed to be harmonic in the circumferential direction, and follows the Bessel function in the radial direction. Bolton and Cao (2015) [72] verified this model by applying harmonic point input excitation both radially and circumferentially. 2.3. Finite element and boundary element models The finite element method (FEM), also known as finite element analysis (FEA), is a numerical technique for solving problems in partial differential form. FEM subdivides the entire body, including the surrounding boundary, into smaller/ simpler parts called finite elements. A system of linear equations that model these finite elements are then integrated to generate the solution equations for the entire body. Unlike the FEM, boundary element method (BEM) subdivides only the boundary of the body into finite elements, and is computationally more efficient. With the advancement of computers, BEM and FEM have become the two dominant numerical techniques in computeraided engineering (CAE). BEM and FEM are often coupled with other methods, and have several derivations, such as waveguide finite element method (WFEM), infinite/finite element method (IFEM), wave finite element (WFE), energy boundary element analysis (EBEA), and energy finite element analysis (EFEA); these methods will be discussed in the following sections. The above methods are often used to calculate the vibrations of the tire, and are used together with fluidstructure interaction (FSI) method to analyze the sound generation and propagation into the air. Sixteen models were reviewed in this subsection. 2.3.1. Toroidal membrane (Saigal et al., 1986) Saigal et al. (1986) [49] modeled the tire as a toroidal membrane under low internal pressure, with the tire mounted on the rim, to investigate its free vibrations. A 12 degree-of-freedom (DOF) membrane finite element was used. To verify the validity of the finite element formulations, the study presented the analytical solutions for three types of models: a flat rectangular membrane, a circular cylindrical membrane, and a toroidal membrane. The final solution of the mode shapes and natural frequencies for the toroidal membrane model is given by
8 h n i XX > ð6f pÞ eiumn t Amn cosðmqÞsin wðq; f; tÞ ¼ > > > 10 > m n < vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ! u pffiffiffi 2 2 > u > > > umn ¼ 1 t m þ 6 2n þ G > : 10 x h
(87)
where w is the displacement normal to the tire surface, q is the coordinate along the toroid circle, f is the coordinate along the meridian circle, and t is the time; m and n are mode numbers, and u is natural frequency; x, h, and G are coefficients given by
8 a2 rh > x¼ > > > Tq > > > > < b h¼ a > > > > > Eh > > > :G ¼ Tq 1 n2
(88)
where a is the radius of the meridian circle, b is the radius of the toroid circle, r is the mass density, h is the thickness of the element; E is Young's modulus, and n is Poisson's ratio; Tq is the internal tension force per unit length in the direction of q; this force equals pa/2, where p is the internal pressure. The above equations are derived under the assumption that b is much greater than a, such that the toroid can be considered as a cylindrical shell of radius a and length 2pb, with the two ends attached to each other. Good agreement with measurements was found for the symmetrical and twisting modes of vibration. It may be noted that this model is restricted only to lower frequencies, as the membrane model cannot consider bending waves at higher frequencies. Hence, it will probably not apply to a steel belted radial tire, but might apply to certain bias ply tires and aircraft Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
23
tires. In addition, the model considers the tire as nonrotating and unloaded. Diaz et al. (2016) [73] developed an analytical tire model that includes all the material behavior for modal analysis under different operating conditions such as speed, load, and inflation pressure. In this study, the effect of Coriolis accelerations was investigated; it was found that real and standing waves are dominant for a nonrotating tire, while complex and travelling waves occur because of speed bifurcation effect. Tire loading generally shifts down the respective modal frequencies. Increasing the inflation pressure increases the structural modal frequencies, but does not influence acoustic modes. 2.3.2. Acoustic radiation (Wright and Koopmann, 1986) The acoustic radiation model (Wright and Koopmann, 1986 [74]) predicted the sound pressure and intensity field around a stationary point-excited tire. First, the surface velocity field of the tire was obtained through conventional modal analysis methods. Then, the sound field was numerically calculated using Helmholtz integral formulation with the information on the tire geometry. The modal analysis was conducted using FEM. The number of elements was 336, which was determined by the wavelength of the structural vibrations being investigated. The wave speeds of the modes are shown in Fig. 10. It can be seen that the wave speed decreases exponentially with increasing mode number. In addition, all the structural wave speeds are well below the speed of sound in air, indicating that in these modes, the tire is an inefficient acoustic radiator, as the tire surface vibrations will not couple very well with the surrounding air. The Helmholtz integral formulation considers the vibrating tire body as a continuous series of monopole and dipole sources. It is shown that the sound intensity field has both positive and negative values, meaning that the tire is absorbing as well as generating sound. This again proves that the non-propagating responses of the tire are very inefficient at radiating the sound. It is suggested that it is the decaying, propagating flexural wave field on the tire that generates the majority of the structure-borne sound. Therefore, the most effective method of reducing structure-borne tire noise is to reduce the forcing inputs at the tire-road interface. 2.3.3. FEA coupling (Richards, 1991) Richards (1991) [61] numerically analyzed the structural-acoustic coupling cavity using FEA. The structural formulation was based on an existing FEA model for various aspects of tire analysis (Richards, 1987 [75]), consisting of isotropic rubber, cord layers, inflation loading, and forces at the fixed nodes. The cylindrical space co-ordinates of each of the nodes are expressed as r, q, and z, while the material coordinate which is attached to the structure is expressed as f. The governing equation is
½K0 þ in½K1 þ n2 ½K2 u2 ½M fag ¼ 0
(89)
where [K0], [K1], and [K2] are the stiffness matrices, [M] is the mass matrix, i is the imaginary unit, n is the integer representing the circumferential wavenumber, u is the natural frequency, {a} is the mode shape containing nodal displacement, element stresses, and the reactions at the fixed nodes. The acoustic formulation is governed by the following wave equation.
V2 j
1 v2 j ¼0 c2 vt 2
(90)
where c is the speed of sound; j is the velocity potential (particle velocity v ¼ Vj), which is related to the acoustic pressure p, by the equation
p ¼ r
vj vt
(91)
Fig. 10. Variation in structural wave speed with frequency for unloaded tire (source from Wright and Koopmann, 1986 [74], Fig. 6; reprinted with permission from Mr. Michael Kaliske of TU Dresden, Germany on behalf of Tire Science and Technology).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
24
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
where r is the gas density. Employing the weight residual method to obtain an approximate solution leads to the following equation.
Z S
vu v2 j E v2 j $dS þ E2 2 20 2 ¼ 0 vt c vt vq
(92)
where S is part of a thin slice [q, q þ dq] on the cavity surface, and u is the displacement of the enclosing structural surface; E0 and E2 are calculated as follows.
8 E0 ¼ > > > < > > > : E2 ¼
Z dV V
Z V
1 r
(93)
dV 2
The resonance frequencies are
sffiffiffiffiffi E un ¼ nc 2 E0
(94)
where √(E2/E0) is considered as the effective radius of the cavity. The modified equation for the structural-acoustic coupling is
½K0 þ in½K1 þ n2 ½K2 u2 ½M fag fRgP ¼ 0
(95)
where {R} includes the complex transformation from f to q, and depends on the circumferential wavenumber n; P is the scaling factor for acoustic pressure. Two additional equations related to these parameters are
8 E > > fRgþ fag 02 P inE0 c0 ¼ 0 > < rc 2 > > > : inE0 P u2 r E0 c0 ¼ 0 E2
(96)
where c0 is the average angular displacement of the gas particles in the cavity, and has the following relationship.
vc v E vj ¼ c einq cos ut ¼ 2 vt vt 0 E 0 vq
(97)
It was found that fluids with high density and pressures may have a significant effect on structural resonances. 2.3.4. 3D two-plate (Larsson and Kropp, 1992) Larsson and Kropp model (1992) [25,26] combines three different modules: (1) a tire model where the dynamic properties of the tire are considered; (2) a contact model where the nonlinear contact forces (both radial and tangential) and the corresponding vibration velocities on the tire surface are calculated; (3) a radiation model where the sound radiation from the tire due to the velocity field on the tire surface is calculated. This is one of the most commonly used TPIN models (Dare, 2012 [13]). It is known that at higher frequencies (above 500 Hz), the curvature of the tire can be neglected, while the importance of the internal structure (i.e., multi-layers of steel and rubber) increases (Larsson and Kropp, 1999 [26]). As a result, Kropp suggested a model of two coupled elastic orthotropic layers/plates (the top layer is the tread, and the bottom layer is the belt) under tension, supported by a linear bedding in the x, y, and z directions, as shown in Fig. 11. The tension in the layer/plate is caused by the internal inflation pressure, and the bedding represents both the side walls and the enclosed air in the tire cavity. The model shows good agreement with measurements up to 2 kHz for a slick 205/60R15 tire (Larsson and Kropp, 1999 [26]). Based on Hamilton's principle (conservation of kinetic and potential energies of a system), the equations of motion for each of the layers are given by Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
25
Fig. 11. Tire model consisting of two coupled elastic layers under tension on an elastic bedding (source from Larsson and Kropp, 1999 [26], Fig. 1; reprinted under fair use provision).
8 " # > > v2 x v2 x v2 x 1 v vx vh vz v2 x v2 x > > > þ þ þ s0 2 ¼ r 2 þ 2þ 2þ G > 2 > 1 2 vx n vx vy vz > vx vy vz vx vt > > > > # > " 2 < v h v2 h v2 h 1 v vx vh vz v2 h v2 h þ þ þ s0 2 ¼ r 2 G þ 2þ 2þ 2 > 1 2 vy n vx vy vz vx vy vz vx vt > > > > > # > " 2 2 2 2 > > v vx vh vz v z v2 z >G v z þ v z þ v z þ 1 > þ þ þ s ¼ r > 0 > : vx2 vy2 vz2 1 2n vz vx vy vz vx2 vt 2
(98)
where G is Young's modulus, and n is Poisson's ratio; x, h, and z are the displacements inside the material in the x, y, and z directions, respectively; s0 is the constant external tension (here, only in the x-direction), r0 is the density of each layer, and t is the time. Modal superposition method is used to calculate the response of a finite structure. However, for high frequencies, a large number of modes is needed, which results in time-consuming calculations and difficulty in modeling the damping. The model was used to compare TPINs on stone mastic asphalt (SMA) and ISO surfaces. TPIN on the SMA surface was predicted to be 2e3 dB louder than that on the ISO surface, while it was 7e8 dB louder in reality [76]. Another study on pass-by TPIN on four different pavements showed accurate results within 5 dB, only in the range of 500e1250 Hz. The model was validated against measured radial and tangential point and transfer mobilities, and good agreement was observed in both magnitude and phase (Andersson and Larsson, 2005 [28]). 2.3.5. TRIAS (de Roo and Gerretsen, 2000) TRIAS (Tyre-Road Interaction Acoustic Simulation, 2000 [77,78]) includes generation mechanism modeling for both structure-borne (tread/texture impact) noise and airborne (air pumping) noise, and sound radiation modeling. The generation mechanism model uses FEM and BEM to predict one-third octave band pressure spectra near the tire. The model includes multiple pavement parameters, such as texture depth, porosity, structure factor, flow resistance, and/or sound absorption, which can be measured experimentally or acquired from the supporting simulation model RODAS (ROad Design Acoustic Simulation), which generates the physical pavement characteristics from the material composition of the mixture (chipping size, binder content, and layer thickness) and analyzes the texture spectrum. The schematic of the full model is shown in Fig. 12. The input pavement parameters include the following: (1) dimensional texture profiles measured in the direction of the road, (2) dimensional texture field, wavelength spectrum, mean profile depth, sound absorption coefficient, distribution of the dimensions of the aggregate particles, binder type and quantity (for porous pavement), thickness of the surface layer, porosity of the surface layer, and specific air flow resistance (de Roo et al., 1999 [78]). Another supporting submodel, TYDAS (TYre Design Acoustic Simulation) generates model inputs from known tire parameters, if all the necessary inputs are not known. The input tire parameters include cross-sectional drawing with dimensions (unloaded and loaded), tread pattern, tread depth, total mass of the tire, density per layer or the average density of the tire, inflation pressure, load, rolling speed, overall stiffness of the tire, and bending stiffness of the combination of tread and carcass in the x (circumferential) and y (axial) directions (de Roo et al., 1999 [78]). The TRIAS model was validated on dense and porous HMA pavements and ISO test tracks. The overall TPIN on dense asphalt pavements was predicted to be within 2 dB, and the TPIN on porous pavements was predicted to be within 5 dB. The model mostly over-predicted the noise. The measured and predicted TPIN spectra have not been presented. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
26
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Fig. 12. Schematic diagram of the mathematical simulation model TRIAS (Tyre-Road Interaction Acoustic Simulation); the supporting model RODAS (ROad Design Acoustic Simulation) is indicated in the upper part of the scheme (source from de Roo and Gerrestsen, 2000 [77], Fig. 1; reprinted under fair use provision).
2.3.6. WFEM (Nilsson, 2004) The application of WFEM to tire vibration analysis was presented first by Nilsson (2004) [79], and later by Fraggstedt (2006) [80], both at KTH Royal Institute of Technology, Sweden. The analysis of the wave field on a stationary tire (no contact with ground) was presented by Sabiniarz and Kropp (2010) [9] at Chalmers University of Technology, Sweden. A waveguide is a structure that guides the waves in a single direction, such as the circumferential direction of the tire, with constant physical properties in that direction. WFEM uses a 2D finite element model [81] over the cross section of the waveguide, and describes the response of the tire in terms of a set of waves in this direction (Sabiniarz and Kropp, 2010 [9]). The response of the nodes in direction X at time t is governed by the following equation.
"
# v2 v v2 K2 2 þ K1 þ K0 þ M 2 VðX; tÞ ¼ FðX; tÞ vX vX vt
(99)
where Ki is the stiffness matrices, M is the mass matrix, F(X, t) is a vector originating from the external load, and V(X, t) contains all the nodal degrees of freedom (DOF). The results of dispersion relations for the waves in the tire are shown in Fig. 13 (Sabiniarz and Kropp, 2010 [9]). There are basically three groups of modes: (1) cross-section modes involving strong out-of-plane motion (Group 1 corresponds to symbols ‘⊥’, ‘’, ‘:’, ‘;’, ‘⋄’, ‘▫’); (2) waves involving significant extensional deformation of the whole cross section (Group 2 corresponds to symbols ‘=’); (3) waves with strong in-plane motion of the sidewalls and weak out-of-plane motion of the
Fig. 13. Dispersion relations with various symbols indicating different groups of modes (Continental 205/55R16, source from Sabiniarz and Kropp, 2010 [9], Fig. 13; reprinted with permission from Elsevier).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
27
belt (Group 3 corresponds to symbols ‘B’, ‘þ’, ‘<’). It was also concluded that Group I has a low wave speed; Group II would be much better radiators; Group III has a higher phase speed, but very low response in the direction normal to the tire surface. 2.3.7. IFEM (Biermann et al., 2007) Biermann et al. (2007) [23] developed a computational tool to predict the sound radiation from rolling tires. The computation process could be divided into three steps: (1) computation of nonlinear tire stationary rolling process; (2) analysis of tire dynamics caused by the road texture; (3) computation of sound radiation using IFEM. The third step employs the structural response calculated by using the first two steps as the boundary conditions. For the sound radiation analysis, the inner region close to and enveloping the tire was discretized with finite elements, while the outer region was discretized with infinite elements. The governing equation for the elements (Astley-Leis elements) is
h i KA þ ðiuÞCA u2 MA p ¼ f A
(100)
where KA, CA, and MA are frequency-independent acoustic system matrices, referred to as inverse mass, admittance, and compressibility operator, respectively; p is the unknown pressure at the nodes of the acoustical mesh, and fA is the load vector representing the excitation of the acoustic fluid due to tire vibration. The model was validated with two tests, a standing tire excited by a shaker and a tire rolling on a drum; a good agreement was found up to 600 Hz. 2.3.8. FEM & BEM (Brinkmeier et al., 2008) Brinkmeier et al. (2008) [17] model is composed of a finite element model of the nonlinear stationary rolling tire and a boundary element model to predict the radiated sound using a modal superposition approach. The overall model uses a combined finite/infinite element approach, where the AstleyeLeis elements method is employed; this ensures a stable and extremely computationally efficient solution process. The finite element model uses the arbitrary Lagrange-Eulerian formulation (ALE). The final equation is
b¼ ðt K WÞD f
tþDt
bf þ t bf t bf e s i
(101)
where tK is the stiffness matrix, and W is the ALE-inertia matrix; Df, with a hat above, is the incremental displacement; tþDtf, with a hat above, is the force resulting from external loads; tfi, with a hat above, is the ALE-inertia; tfs, with a hat above, is the internal stress. The equation for the boundary element model is
i h AðuÞp :¼ KA þ ðiuÞCA u2 MA p ¼ f A
(102)
where KA is the inverse mass operator or acoustic stiffness, CA is the admittance operator or acoustic damping, MA is the compressibility operator or acoustic mass matrix, p is the sound pressure, and fA is the load vector representing the tire dynamics on the surface. The symbol “:¼” indicates that it is a definition. The model successfully predicted the acoustic ranking of tires, but the error between the predicted and experimental results was in the range of 6e17 dB, probably because the model considers only the structural-borne noise at low frequencies (100e850 Hz), and ignores pavement absorption. In addition, only vertical contact forces are considered, and friction forces are not included. 2.3.9. WFE (Waki et al., 2011) Waki et al. (2011) [82] applied WFE method (Waki et al., 2009 [83]) for primitive estimation of tire vibration and noise radiation. The study provided insight into the locations of the radiating parts on the tire. In contrast to FEA, which could investigate vibrations only at low frequencies, WFE could predict vibrations up to high frequencies around 1 kHz. A short section of a tire, having a uniform structure, was modeled using FEM (Díaz et al., 2012 [84]). The equation of motion is
Dq ¼ f D ¼ K þ juC u2 M
(103)
where D is the dynamic stiffness matrix, q is the nodal displacement vector, and f is the nodal force vector; K, C, and M are the stiffness, damping matrix, and mass matrix, respectively. Then, the periodicity condition, as shown below, was applied to the equations of motion of the short section to formulate the eigenproblem.
qR ¼ lqL f R ¼ lf L
(104)
where the superscripts L and R indicate the left- and right-hand sides, respectively, of the section. l is related to the wavenumber k as given below.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
28
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
li ¼ expðjki DÞ
(105)
where D is the length of the short section. It was found that the sound emission field could be very different from the vibration field. 2.3.10. Road input (Kido et al., 2011) Kido et al. (2011) [85] correlated the actual road profile with excitation at the tire contact patch. The spatial road profile was transformed into time domain profile, considering the driving speed, and was then transformed into frequency domain data. The profile data were two-dimensional, which means that the entire two-dimensional tread area could be considered. Further, an FE model was developed to correlate the tire patch excitation and the axle forces under steady-state conditions. Both Coriolis force and centrifugal force due to tire rotation were considered. The response of the wheel assembly x, excited by the external force f with tire rotational angular velocity of ur (rad/s), is represented by the following equation.
Mx€ þ ðC þ ur M1 Þx_ þ K u2r M2 x ¼ f
(106)
where M, K, and C are the mass matrix, stiffness matrix, and viscous damping matrix, respectively, based on the tire vibrational properties; M1 is the skew-symmetric matrix, and M2 is the symmetric matrix that depends on the inertia property and rotational axis of the tire. The Coriolis force is in proportion to the velocity urM1ẋ, and the centrifugal force is in proportion to the displacement u2r M2x. The model also considers the tire stiffness drop caused by tire rotation. The calculations of the longitudinal/lateral forces on the axle, and the accelerations at low frequencies (0e120 Hz) were found to have good agreement with experimental measurements. It was also found that the influences of Coriolis force and centrifugal force are especially important in the frequency range of 80e100 Hz. 2.3.11. FSI (Yang et al., 2013) The FSI model developed by Yang et al. (2013) [22] consists of two submodels: the rolling tire-pavement interaction submodel and the near-field sound propagation submodel. The two submodels are coupled using the FSI method, where the equilibrium conditions of displacement and stress are fulfilled. The tire-pavement interaction is assumed to be a two-body contact problem, where the pavement surface is assumed to be rigid, and the body forces are ignored. The relevant equation is given below.
Z t
V
rT u€T $duT dt V þ
Z t
V
cT u_ T $duT dt V þ
Z t t
V
Z
tT deT dt V ¼ t
Sf
t S f T $duS dt V
Z t
t
f C $duT dt S
(107)
Sc
where T represents the tire, rT is the tire density, cT is the tire damping ratio, ttT is the stress tensor, deT is the strain corresponding to the virtual displacement duT, tfTS is the surface force per unit area, and tfC is the surface force per unit area on contact surface. The near-field sound propagation submodel is described using the large eddy simulation (LES) model (Wagner et al., 2007 [86]). The governing equations in the LES model for an arbitrary moving cell with a volume V(t) and a cell-face area S(t) are given by
d ∭ rdV þ ∬ SðtÞ rðU WÞ$dS ¼ 0 dt VðtÞ
(108)
d ∭ rUdV þ ∬ SðtÞ ½rUðU WÞ s$dS ¼ ∭ VðtÞ f B dV dt VðtÞ
(109)
where r is the air density, U is the flow velocity vector, W is the moving mesh velocity vector, fB is the air body force, and s is the total stress tensor. This model focuses only on the PIARC passenger car tire (PIARC, Permanent International Association of Road Congresses) (Sandberg, 1990 [87]; PIARC, 2004 [88]). The A-weighted one-third octave band sound pressure level showed good agreement between simulation and experimental results in the range of 500e2500 Hz, with vehicle speed in the range of 50e90 km/h. 2.3.12. Wear (Tong et al., 2013) The wear model developed by Tong et al. (2013) [89] is composed of a steady rolling tire model, a tire wear model, and a tire acoustic model. The steady rolling tire model used the ALE approach, by decomposing the motion into a pure rigid body motion and superimposing the deformation. The expressions for the velocity and acceleration of the particle in both Lagrangian and Eulerian reference frame were derived.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
29
The tire wear model included an FE model of the 3D tread pattern prepared using Pro/E and Abaqus FE software. The tire acoustic model perceived the sound field as an ideal homogeneous fluid medium with small-amplitude waves. The acoustic excitation was obtained from steady rolling tire model. A modal test with random signal excitation was performed to compare the measured modal frequency with the simulation results. The results showed good agreement with a maximum deviation of 4.48% in the frequency range of 0e500 Hz. The model investigated four factors that potentially influenced TPIN: speed, tire load, inflation pressure, and tire wear. 2.3.13. EFEA/EBEA (Vlahopoulos et al., 2013) Vlahopoulos et al. (2013) [90] investigated the airborne noise inside a vehicle, originating from exterior components such as tires, engine, and exhaust. These exterior components were assumed to be the acoustic source, and then EBEA was applied to compute the exterior acoustic loading on the vehicle. Further, EFEA was employed to calculate the interior airborne noise with the exterior acoustic loading as the excitation, i.e., EFEA dealt with acoustic transmission through the vehicle mass panel, constraint layer damping (CLD), and acoustic absorption treatment (or noise reduction) from the exterior to the interior. The power transfer mechanism is accounted for by the mass barrier effect of the panel. The governing equation is given below.
2
tnormal ¼
3
ri ci 6 4 rj cj
4
7
2 2 5 B 1 þ rri ccij þ u rs trs þm c j j j
(110)
where tnormal is the normal mass barrier transmission coefficient from acoustic enclosure i to acoustic enclosure j through the panel; ri, rj, and rs are the densities of the acoustic enclosure i, acoustic enclosure j, and panel s, respectively; ci and cj represent the speed of sound in acoustic enclosures i and j; ts is the thickness of the panel separating the two acoustic spaces; u is the angular frequency, and mB is the surface mass density of the mass barrier. It is known that FEA typically models only structure-borne noise at low frequencies, and a detailed FE model is needed in the path between the excitation and the receiver. As compared to FEA, EFEA captures the physics associated with the path between the exterior acoustic excitation and the interior acoustic cavity, and covers the entire frequency range of 200e8000 Hz. In this vehicle model, it was found in the contribution analysis, expressed in terms of the power flow into the interior acoustic space of the vehicle, that the front floor contributed the most, up to 70.8%. 2.3.14. Cavity (Mohamed and Wang, 2015) The cavity model developed by Mohamed and Wang (2015) [91] focuses on the noise generated from the tire torus cavity resonance. The tread plate is assumed to be flexible, while the other walls are considered as rigid. A modal series solution of the sound pressure frequency response inside the tire cavity is derived from the wave equation using modal superposition. This work also provides a noise reduction method, which involves placing a trim layer on to the inner surface of the tire tread plate by adding a damping loss term in the sound pressure frequency response function (sound pressure peak reduction is 10e15 dB). The cavity natural frequencies can be approximated by the following equation.
fmnl
c ¼ 2p
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2 2 lp 2np 4m þ þ Do þ Di W Do Di
(111)
where fmnl is the cavity natural frequency at the mode corresponding to the indexes m, n, and l in the axial, radial, and azimuthal direction, respectively; c is the speed of sound, W is the tire width, Do is the outer diameter of the tire torus, Di is the inner diameter of the tire torus. The calculated results are also compared with FEM results. It may be noted that only a minimum of six elements per wavelength is required in the acoustic space, and hence FEM involves relatively less computation. The Donnel shell theory [92] is used for the calculation of tire tread vibration. The tread surface particle velocity is defined by
upr ¼ iu
F0 eiut r ht
∞ X B2ml sinðlpz0 =WÞcosðmq0 Þsinðlpz=WÞcosðmqÞ u2lm 1 þ ihp u2 m¼0;l¼0
(112)
where upr is the solid particle velocity of the plate surface in the r direction, r is the mass density of the top tire tread pattern plate, ht is the plate thickness, m and l are the plate wave numbers in the q and z directions, respectively; q0 and z0 are the location coordinates of the harmonic point excitation force, F0 is the amplitude of the harmonic point excitation force, ulm is the natural circular frequency of the annular outer plate, hp is the plate damping loss factor, and Bml is the normalized plate mode shape coefficient.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
30
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
2.3.15. MLE (Wei et al., 2016) Wei et al. (2016) [93] developed a model using mixed Lagrangian-Eulerian (MLE) method to predict the tire noise in the time domain. It was pointed out that the tire vibrations must be analyzed in the Lagrangian reference, while the acoustic analysis is normally performed in the Eulerian reference. The MLE model transformed the tire vibration data (velocity/acceleration) in the Lagrangian mesh to the Eulerian mesh (finite element model). The radiated noise was then calculated from the information on the vibration of the tire outer surface (tread pattern and groove surface) using acoustic BEM. It was found that the accelerations are largely distributed near the footprint (approximately double the contact length). The accelerations at the trailing edge are lower than those at the leading edge, but with a much wider distribution, as shown in Fig. 14; this explains the reason for the noise near the trailing edge to be greater than at the leading edge. The model considered only the tire noise due to vibrations, such as impact between the tread pattern and the road texture, and did not include air pumping noise. 2.3.16. Time domain BEM (Banz et al., 2016) Time model (Banz et al., 2016) [94] used a time domain Galerkin boundary element method based on a time-dependent integral equation of the second kind to formulate the acoustic wave equation for tire noise radiation with the tire dynamics given. The governing equation to be solved is given below.
ðI þ K 0 Þ4ðt; xÞ ¼ 2
vp v2 un vpI ðt; xÞ ¼ 2rc2 2 ðt; xÞ 2 ðt; xÞ vn vn vt
(113)
where I þ K0 is an operator that accounts for the partially absorbing, acoustic boundary conditions on the road, 4 is the unknown density, t is the scaled time given by t ¼ ct, in which c is the speed of sound and t is the real time; x is the 3D spatial coordinate vector; p is the sound pressure, n is the direction normal to the tire surface; r is the air density, un is the displacement of the tire in the outer normal direction, and pI denotes the incident wave. The model was validated in three applications: (1) the sound emission of steady-state vibrating tires (205/55R16 passenger car tire and 315/80R22.5 rib truck tire); (2) noise amplification in the horn-like geometry between the road and the tire (205/55R16 slick tire); (3) Doppler effect (2% shift) of a moving tire (artificial signal). 2.4. Computational fluid dynamics models CFD uses numerical algorithms to solve and analyze fluid flow problems in computer-aided engineering (CAE). It can be applied to the modeling of airborne TPIN. Five models are reviewed in this subsection. 2.4.1. Shock wave (Gagen, 1999) Gagen (1999 [95], 2000 [96]) modeled the acoustic sources of air pumping from the squeezed tread groove cavities using the shock wave theory (Kinsler et al., 1999 [97]) instead of the commonly used monopole theory (Hayden, 1971 [29]). It was pointed out that at high speeds, as the tire/pavement impact causes the groove walls to move with velocities of the order of 1 m/s and accelerations of 103 m/s2 for a period more than 1 ms, the groove volume decreases (and then increases at the trailing edge) rapidly by around 10%, which is much faster than the rate at which air can evacuate from or suck into the squeezed/expanded groove. This leads to significant pressure and density fluctuations, which makes the linear approximations (small amplitude monopole theory) invalid. As per the nonlinear acoustic phenomenon, both thin-width grooves and large-width grooves generate low noise, while medium-width grooves generate large noise (Ejsmont et al., 1984 [98]).
Fig. 14. Left - Footprint and contact pressure distribution at 70% load (source from Wei et al., 2016 [93], Fig. 6; reprinted with permission from Mr. Michael Kaliske of TU Dresden, Germany on behalf of Tire Science and Technology); right - Acceleration at the footprint at 0.0024 s (source from Wei et al., 2016 [93], Fig. 7; reprinted with permission from Mr. Michael Kaliske of TU Dresden, Germany on behalf of Tire Science and Technology).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
31
The monopole equation for a simple harmonic source is expressed as
grV€ gru2 V ¼ 4pr 4pr
p1 ðr; tÞz
(114)
where p1 is the sound pressure, r is the distance between the sound source and receiver, t is the time, g is the ratio of specific heats (1.4 for air), r is the air density, V is the air volume, V with double dots is the second order derivative of the air volume with respect to time, u is the frequency of sound. All the parameters are normalized to be dimensionless, so that the speed of sound is unity. For non-simple harmonic source, Fourier decomposition into frequency components should be performed before applying the monopole theory. The shock wave equation for propagation of density disturbance is given by
0 ¼ vtt r
vxx rg
g
_c € v r _ f c t
f_vc rvy v r v v v r v v vcc rg f 2f vt r cx x y cc y c 2 v þ v r v r v v r þ xx x c x c x f f f f f gf 2 f2 (115)
where v is the derivative symbol, vx and vy are the fluid velocities in the x and y coordinates, respectively; c is the computational coordinate where the groove walls are stationary, and is given by y ¼ f(t)c and vc ¼ (vy ḟc)/f. In this study, the kinetic energy of the emitted jet was also investigated to obtain the sound intensity in the far field. 2.4.2. Hybrid CFD (Kim et al., 2006) The hybrid CFD model developed by Kim et al. (2006) [99] consisted of three submodels: (1) small-scale noise generation submodel; (2) noise propagation submodel; (3) far-field acoustic pressure prediction submodel. In the first submodel, the air-pumping noise source was simulated as piston/sliding-door/cavity geometry, as shown in Fig. 15 (Kim et al., 2006 [99]), using Navier-Stokes finite volume solver in CFD, which considers the motion of the road surface relative to the tire. The second submodel transferred the flow properties in the tire grooves, obtained in the first submodel, as nonlinear airpumping sources to the flow simulation of the full tire/road domain considering mid-scale acoustic scattering process. Finally, the third submodel predicted the far-field acoustic pressure by the linear Kirchhoff integral method with the flow properties around the rolling tire and road surface geometry, acquired from CFD calculation. The linear Kirchhoff integral equation is given below.
p0 ðx; tÞ ¼
1 4p
Z Z
cos q 0 1 vp0 cos q vp0 þ dSðy; tÞ p r vn cr vt r2
(116)
S
where p0 is the perturbed pressure, x and t are the observer's location vector and time respectively; y and t are the source location vector and retarded time variables, respectively; q is the angle between the normal vector (n) on the surface and the radiation vector (r), and r is the distance between an observer and a source at the retarded time. The pressure and its derivatives are calculated at the retarded time. It was found that TPIN at frequencies in the range of 3e8 kHz was suppressed by the ventilation of the tire groove.
Fig. 15. Schematic diagrams of the piston/sliding-door/cavity source model (source from Kim et al., 2006 [99], Fig. 3; reprinted with permission from AIP Publishing LLC).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
32
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
2.4.3. CFD (Chen et al., 2014) Chen et al. (2014) [11] used CFD to investigate airborne TPIN, such as air pumping, air resonant radiation, and pipe resonance. Different tread patterns (hand carved) were simulated and compared with measured near-field noise using the laboratory drum method, and the deviation was found to be a maximum of 4.1% in the frequency range of 0e2000 Hz. 2.4.4. Air pumping CFD (Gautam and Chandy, 2016) Gautam and Chandy (2016) [100] investigated the air pumping noise generation due to the compression and expansion of air in the tread grooves using an approach based on LES. The compression and expansion of air was assumed to result from the change in the groove volume. The study suggested two types of tire groove deformation that resulted in the change in the groove volume: (1) a commonly used piston-type motion of the bottom wall and (2) a more realistic bulging in/out of the side walls of a tire groove, as shown in Fig. 16. In this figure, the white box refers to the groove, and the black box refers to the road. Numerical CFD analysis was performed using ANSYS FLUENT™ for both types of deformation. It was found that the piston deformations result in a distinct frequency peak around 2e4 kHz, while the side wall deformations result in a more broadband noise spectrum. However, the study assumed that the expansion period begins immediately after the compression period, which is not the case in reality. A further study by Gautam and Chandy (2016) [101] investigated both the near-field and far-field acoustic characteristics using LES and Ffowcs-Williams and Hawkings (FW-H) approach, respectively, considering several airborne mechanisms, such as air pumping, pipe resonance, Helmholtz resonance, and rotational turbulence. The simulation environment was a slick tire (215/ 60R16), with two transverse slots closely spaced in the tread, rotating on a smooth drum at 40 km/h in a semi-anechoic chamber. The change in the slot volume was assumed to be 22% (Takami and Furukawa, 2015 [102,103]) at the tire-pavement interface. It was concluded that the air pumping noise is focused at higher frequencies around 3e5 kHz, especially at the trailing edge of the contact patch. The horn amplification effect can be up to 20 dBA starting from 1.6 kHz, which is proved by the acoustic directivity analysis. However, validation experiments corresponding to the same conditions as in the case of simulation were not reported. 2.4.5. Pipe resonance (Fabrizi, 2016) In this study (Fabrizi, 2016 [104]), the pipe resonance mechanism was modeled using CFD and computational aeroacoustics. Only the aeroacoustic noise-generation mechanism of the pipe resonance phenomenon was covered; i.e., this is different from the amplification mechanism of pipe resonance, where vibroacoustic waves are also considered. The analysis of aeroacoustics in the longitudinal tread grooved is based on Lighthill acoustic analogy (Lighthill, 1952 [105]; Lighthill, 1954 [106]).
1 v2 p0 c20
v2 p0
vfi v2
v2 ¼ r u u t þ i j ij vxi vxj vxi vt 2 vt 2 vx2i
p0 ¼ p p0
p0 c20
! 0
r
(117)
r0 ¼ r r0 where c0 is the speed of sound, t is the time, p and p0 are the instantaneous and equilibrium air pressures, r and r0 are the instantaneous and equilibrium air densities, ui and uj are air velocities in the i direction (flow direction) and j direction (perpendicular to flow direction), respectively; tij is the shear stress tensor, fi is the fluctuating external force per unit volume.
Fig. 16. Schematic diagram showing different deformation models with (a) piston-like motion of bottom wall (b) side walls bulging inward (source from Gautam and Chandy, 2016 [100], Fig. 3; reprinted with permission from Ms. Beth Darchi on behalf of ASME).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
33
Assuming that the flow can produce sound only at high velocities, the following equation is obtained.
vTij ðy; tÞ 1 xi xj ∭ dVy 4pc40 kxk3 vt2 Tij ¼ rui uj þ p rc20 dij tij
rðx; tÞz
(118)
where t is the time, x is the observer point (xi, xj), y is the source point, Tij is the Lighthill stress tensor or quadrupole strength density, and dij is the compressive stress tensor coefficient. The tire model used in the aerodynamic analysis is of size 205/55R16 with two longitudinal tread grooves, with a rotating speed of 80 km/h. Through LES, three phenomena related to the tire aerodynamic noise have been identified: (1) At the pipe inlet, the air flow is squeezed laterally into the groove from both sides, each with a high swirl component (counter-rotating vortices); (2) at the pipe outlet, the flow creates jets that develop inside the recirculation bubble of the wake; (3) the flow accelerates laterally to a speed of around 300 km/h and detaches, generating a high-energy shear layer. The sound pressures at four positions were analyzed; the results are shown in Fig. 17. Only for the positions of the pipe and wake, a clear resonance peak was observed at 1700 Hz, which matches the theory, if the correction length is set to 10 mm (two times the pipe diameter). This suggests that the resonant acoustic wave could propagate (acoustic energy) only in the direction of the flow (fluid energy). This directivity is similar to that of a dipole oriented in the longitudinal direction. The peak at 2100 Hz for the positions on the side of the leading edge is possibly the effect of the jet flow generated at the shoulder rib by the lateral squeezing of air out of the rolling tire. It may be noted that a peak around 600e800 Hz occurs for all positions, which is likely to be the result of the vibrations of the tire structure. 2.5. Summary The conventional physics models are summarized in Table 4. It can be seen that these models have been continuously improved since 1971. The tire was modeled as a ring, shell, or plate for vibration investigation; the tire/rim assembly was modeled as a structure coupled with an air cavity; the sound source was modeled as a monopole. However, most models focused only on structure-borne noise, especially the noise due to the impact between the tire and pavement. Therefore, the tire vibration was the output parameter instead of sound pressure level. The frequency range was typical below 500 Hz; some models claimed to be able to predict higher frequencies, but still they were limited to structure-borne noise. In addition, many of the models considered only a stationary tire, which is quite different from the normal driving conditions. Only one specific tire or one specific pavement was investigated in some of the models. The finite element and boundary element models are summarized in Table 5. FEM has been applied to TPIN modeling since 1986. Similar to the conventional physics model, the finite element and boundary element models also generally focused on structure-borne noise, and considered the tire carcass as a membrane or plate. However, the near-field or far-field sound pressure level could be easily predicted with the structure-fluid coupling technique. The frequency range was almost exclusively below 500 Hz or even lower, such as 150 Hz, because of the inherent drawback of the FEM. Vehicle speed was typically included in the modeling as an input. Tire parameters were investigated in detail, but not much attention has been paid to the pavement parameters. The CFD models are summarized in Table 6. The use of CFD models started in 1999, i.e., later than the previous two categories. With the help of CFD, the TPIN modeling was extended from structure-borne noise to include airborne noise (air
Fig. 17. On the left, A-weighted sound pressure levels of the pressure signals recorded with four probes placed around and inside the contact patch of the tire with the road surface; on the right, the position of the four probes with the reference of the footprint shape (source from Fabrizi, 2016 [104], Fig. 12; reprinted with permission from Mr. Michael Kaliske of TU Dresden, Germany on behalf of Tire Science and Technology).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
34
No. Model name
Author
Year
1
Monopole
Hayden
1971 [29] Monopole
2
Unified
Plotkin and Stusnick 1981 [30] Thin shell
3
Truck tire
Keltie
1982 [38] Thin shell
4 5
2D ring Various mechanisms Tire vibration
Kung et al. Heckl
1986 [39] Modal expansion 1986 [44] Various
Near-field, normalized radiation 1/3 octave band sound power level Near-field, SPL LDV, vibration, SPL
Bremner et al.
1997 [46] SEA, wavenumber decomposition 1997 [51] Image source
Far-field (7.5 m), 1/3 octave band SPL Impedance, absorption
1998 [52] Wavenumber decomposition
6
rengier et al. Be
Ref. Method
Output
Frequency Mechanism range [Hz]
Sound pressure, pure tone N/A
Air pumping
0e8000
Driver parameter Tire parameter
Pavement Environment parameter parameter
Speed
N/A
N/A
Texture
N/A
Impact, air pumping 100e2000 Structure-borne
Speed, tire load
0e400 0e2000
Tire load Rotating, loaded
Structure-borne Various
Inflation
Groove geometry Various
Damping, N/A N/A bending stiffness Various Texture N/A Tread pattern Roughness N/A
100e4000 Impact
Speed, 40 psi
Tread pattern, materials N/A
Texture
N/A
100e5000 Pavement absorption
N/A
Point mobility, wavenumber
0e1000
215/70R14
N/A
N/A Rough, 1 30 mm step N/A N/A
100/50R12
N/A
N/A
Geometry, structure
N/A
N/A
N/A Airflow resistivity, porosity, tortuosity N/A N/A
7
Porous pavement
8
Wavenumber Bolton et al. decomposition
9
Orthotropic plate
Kropp
1999 [53] Multipole synthesis
Vibration, 1/3 octave band SPL
125e2500 Structure-borne
Stationary, inflation, gas type Unloaded
10
Shell
Kim and Bolton
Vibration
0e4000
Point impact
Speed
11
Molisani et al. Kim et al.
Modal amplitude, force at spindle Vibration, flexural wave, eigen-frequency
0e1250
12
Coupled cavity Waveguide
0e2000
Cavity resonance Stationary without rotating Structure-borne Stationary
13
Cavity
Feng et al.
Cavity resonance Speed, tire load
N/A
N/A
Temperature
3D multilayer
O'Boy and Dowling
Cavity SPL, narrowband Vibration
100e300
14
2002 [57] Wave number decomposition 2003 [60] Donnell-Mushtari theory, structural acoustic 2007 [62] Spring-mass, elastically supported beam, cylindrical shell 2009 [63] Coordinate transformation 2009 [65] Bessel equation
0e2000
Structure-borne
225/45ZR17
N/A
20 C
15
Green's function Coupled mode
Arteaga
2011 [68] Modal-ALE
Vibration
0e300
Impact
185/70SR14
N/A
N/A
Bolton and Cao
2013 [70] Structural-acoustic coupling
LDV, vibration
0e6000
Structure-borne
225/55R17
N/A
N/A
16
Structure-borne
Tire load, break/acc. Tire load, speed Stationary & 50 m/s, 3 bar
Slick
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 4 Summary of conventional physics models on TPIN.
No.
Model name
Author
Year
Ref.
Method
Output
Frequency range [Hz]
Mechanism
Driver parameter
Tire parameter
Pavement parameter
Environment parameter
1
Toroidal membrane
Saigal et al.
1986
[49]
Membrane FE
Vibration
Low
Structure-borne
Geometry, structure
N/A
N/A
2
Wright and Koopmann Richards
1986
[74]
N/A
N/A
N/A
[61]
Near-field, sound pressure, intensity Vibration
Vibration, radiation
1991
0e400
Cavity resonance
185/70R14
N/A
N/A
4
3D two-plate
1992
[25]
Pass-by, SPL
0e2000
Structure-borne
Tire load
N/A
TRIAS
2000
[77]
FEA, BEM
N/A
Impact, air pumping
N/A
WFEM
2004
[79]
WFEM
0e1500
Structure-borne
Stationary
Geometry, structure
Various, RODAS N/A
N/A
6
N/A
7
IFEM
Biermann et al.
2007
[23]
0e1400
Impact
Speed
Slick
Texture
N/A
8
FEM & BEM
2008
[17]
Structure-borne
40-120 km/h
N/A
2011
[82]
0e1500
Impact
10 11
Road input FSI
Kido et al. Yang et al.
2011 2013
[85] [22]
FE, modal decomposition FE FEA
0e150 500e2500
Texture impact TPIN
225/45R17 265/70R17 Patch area PIARC
Texture N/A
N/A N/A
12
Wear
Tong et al.
2013
[89]
Acoustic FE
Axle vibration Near-field, 1/3 octave band SPL Near field, SPL
Inflation: 220/180 kPa Speed, tire load 70-90 km/h
Texture spectrum N/A
N/A
WFE
Pass-by, 1/3 octave band SPL Vibration
100e850
9
Brinkmeier et al. Waki et al.
IFEM, Astley-Leis elements FEA, BEM, ALE
Near field, 1/3 octave band SPL; pass-by, SPL Point mobility, wavenumber Far-field (1 m), 1/3 octave band SPL
Various, slick 205/60R15 Various, TYDAS
Texture
5
Larsson and Kropp De Roo and Gerretsen Nilsson
FE, Helmholtz integral Weighted residual approximation FE
80e200
3
Acoustic radiation FEA coupling
Inflation, nonrotating, unloaded Stationary, point-excited N/A
0e500
Impact
Wear
N/A
N/A
13
EFEA/EBEA
Vlahopoulos et al.
2013
[90]
Interior noise
200e8000
Airborne
N/A
N/A
N/A
14
Cavity
2015
[91]
Tire cavity SPL
150e500
Torus cavity resonance
N/A
Tire size
N/A
N/A
15
MLE
Mohamed and Wang Wei et al.
2016
[93]
Near field, 1/3 octave band SPL
Time domain, 0e1250
Vibration
70 km/h
N/A
Time domain BEM
Banz et al.
2016
[94]
1/3 octave band SPL
200e2000
Radiation
Steady-state moving
Truck/bus, 315/60R22.5 385/65R22.5 Passenger/truck, slick
N/A
16
EFEA, EBEA, panel contribution analysis FEM, Bessel root method Mixed Lagrangian-Eulerian, FE, BEM BEM
Tire load, speed, inflation N/A
N/A
N/A
N/A
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89 35
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 5 Summary of finite element and boundary element models on TPIN.
36
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Table 6 Summary of computational fluid dynamics models on TPIN. No. Model name
Author
Year Ref.
Method
1
Gagen
1999 [95]
CFD
2
3 4
5
Shock wave Hybrid CFD
Output
Frequency Mechanism range [Hz]
Near-field, sound N/A, time Air pumping pressure/energy domain Far-field, 1/3 250e8000 Air pumping, Kim et al. 2006 [99] CFD, octave band SPL pipe Kirchhoff resonance integral CFD Chen 2014 [11] CFD Near-field, 1/3 0e2000 Airborne et al. octave band SPL 1000 Airborne, air Near-/far-field, Gautam 2016 [100] CFD Air e10000 pumping drum, pumping and narrowband SPL Chandy CFD 1500 Pipe Pipe Fabrizi 2016 [104] CFD, LES, Pass-by/ine2000 resonance resonance aeroacoustics groove, narrowband SPL
Driver Tire parameter Pavement Environment parameter parameter parameter Speed
Groove angle, geometry 100 km/h Ventilation groove, 165/ 70SR13 N/A Tread pattern 40 km/h
80 km/h
215/60R16, transverse slots 205/55R16, longitudinal grooves
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
Drum with N/A Safety Walk
pumping mechanism). Near-field sound pressure level was widely accepted as the output parameter of the model. The tire tread pattern was the most important input parameter investigated. However, none of the current models included pavement parameters; the authors believe that these are also important, because air voids are formed within the contact patch between the tread pattern and the pavement texture.
3. Statistical models 3.1. Introduction As the physics of TPIN generation is uncovered to a great extent, statistical models are more common, especially in the industry. Statistical or semi-empirical models are based on correlation of measured noise data for various tire-pavement parameters. For example, the measured TPIN can be correlated with pavement texture, including texture levels at different wavelengths in the longitudinal direction (Rasmussen et al., 2006 [107]) and overall texture mean profile depth (MPD) (Fujikawa et al., 2009 [108]). The early models, shown in Tables 7 and 8, were used to predict pass-by traffic noise. They used L50 as the indicator during the period between the 1950's and 1960's, and then Leq prevailed as the indicator (Quartieri et al., 2009 [109]). All these models essentially have the following format (Murphy and King, 2014 [118]).
Lp ¼ E Atot þ C
(119)
where Lp is the noise level to be predicted, E is the emission of the source (sound power of a single point source, or sound power per unit length of a line source), Atot is the total sound attenuation (geometric divergence, atmospheric absorption, ground absorption, the barrier, and miscellaneous effects), and C is the correction factor (façade reflection, different pavement/vehicle types). Lp can be in the form of L10,18h, LAeq or Lden. L10,18h is the noise level just exceeded for 10% of the 18-h measurement period from 06:00 to 24:00 h. LAeq is the A-weighted equivalent sound level. Lden is the day-evening-night equivalent level (A-weighted equivalent sound level), measured over the 24-h period, with a 10-dB penalty added to the levels between 23:00 and 07:00 h and a 5-dB penalty added to the levels between 19:00 and 23:00 h to reflect people's extra sensitivity to noise during the night and the evening. These statistical models have better accuracy than deterministic models, as the model parameter values are obtained from the actual measured data. However, empirical measurements can never account for all potential conditions, and hence
Table 7 Nomenclature of early models used for traffic noise prediction (Quartieri et al., 2009 [109]). Symbol
Meaning
Unit
LXX Leq Q Qeq d v P n
Noise level exceeded for XX% of the measurement period Equivalent (energy) sound level Traffic volume (QM for motorcycles, QBUS for buses) Equivalent vehicular flows Distance from the traffic lane center to the receiver Vehicle speed Percentage of heavy vehicles in the traffic Acoustical equivalent of heavy vehicles, indicating the number of light vehicles that generate the same acoustic energy of a heavy vehicle
[dB] or [dBA] [dB] or [dBA] [Vehicles per hour] [Vehicles per hour] [ft] [mph] [%] []
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
37
Table 8 Early models used for traffic noise prediction (Quartieri et al., 2009 [109]). No.
Reference
Author
Year
Equation
1
[110]
Anonymous
1952
2
[111]
Nickson
1965
3 4
[112] [113]
Johnson and Saunders Griffiths and Langdon
1968 1968
5 6 7
[114] [115] [116]
1969 1977 1991
8
[117]
Galloway Burgess Centre Scientifique et Technique du Batiment (France) Fagotti and Poggi
L50 ¼ 68 þ 8.5log10(Q) 20log10(d), for speed of 55e75 km/h and d > 20 L50 ¼ C þ 10log10(Q/d), where C is a constant that is determined from experimental data under given traffic conditions (vehicle and pavement types) L50 ¼ 3.5 þ 10log10(Qv3/d) Leq ¼ L50 þ 0.018(L10 L90)2 L10 ¼ 61 þ 8.4log10(Q) þ 0.15P 11.5log10(d) L50 ¼ 44.8 þ 10.8log10(Q) þ 0.12P 9.6log10(d) L10 ¼ 39.1 þ 10.5log10(Q) þ 0.06P 9.3log10(d) L50 ¼ 20 þ 10log10(Qv2/d) þ 0.4P Leq ¼ 55.5 þ 10.2log10(Q) þ 0.3P 19.3log10(d) Qeq ¼ Q[1 þ P(n 1)/100] L50 ¼ 11.9log10(Qeq) þ 31.4 Leq ¼ 0.65L50 þ 28.8 Leq ¼ 10log10(QL þ QM þ 8QP þ 88QBUS) þ33.5
1995
statistical confidence level and confidence interval are normally analyzed. In addition, some of the parameters in these models are not real physical parameters, but derived parameters such as a texture spectrum [14], which means that they are not adjustable directly. To change these parameters, they have to be translated back to the original physical parameters, which are not always explicit or even impossible for nonlinear relationships. Various statistical analysis algorithms can be used. In this section, 18 statistical models are reviewed; they consist of 10 traditional regression models, 3 principal component analysis models, and 5 fuzzy curve-fitting models. The model name, method, output noise parameter with noise measurement technique, noise generation mechanisms investigated, and the input parameters of each model are listed and compared in the tables as a summary in section 3.5. The summary tables can be used as a reference to direct the readers to the models of their interest. 3.2. Traditional regression models This subsection covers the traditional regression models, which include linear regression models and nonlinear regression models. The linear regression is easier to model, while the nonlinear regression is more flexible. 3.2.1. Germany (RLS, 1990) €rmschutz an Strassen, or Guideline for noise protection on streets) was The first version of RLS (Richtlinien für den La released by Der Bundesminister für Verkehr (now BMVI, Federal Ministry of Transport and digital Infrastructure) in 1981 [119]. The current version is RLS 90 [120], released in 1990. It is used to predict the sound level at a distance of 25 m from the center of the traffic lane. The nomenclature for this model is presented in Table 9 (RLS, 1990 [120]). First, the average level (reference value) is calculated as ð25Þ
Lm;E ¼ 37:7 þ 10 log10 ðQ ð1 þ 0:082PÞÞ
(120)
Accounting for the vehicle and pavement conditions, the real average sound level at 25 m distance is given by ð25Þ
Lm ¼ Lm;E þ RSL þ RRS þ RRF þ RE þ RDA þ RGA þ RTB
(121)
where RRS ¼ 0.6jgj 3 (set to 0 if negative). RSL is given by
RSL ¼ LPkw 37:3 þ 10 log10
! 100 þ 100:1D 1 P 100 þ 8:23P
(122)
where
8 < LLkw ¼ 23:1 þ 12:5log10 h ðVLkw Þ i LPkw ¼ 27:7 þ 10log10 1 þ ð0:02vPkw Þ3 : D ¼ LPkw LLkw
(123)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
38
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Table 9 Nomenclature for RLS 90 model (RLS, 1990 [120]). Symbol
Meaning
Unit
L(25) m,E Lm Q P RSL RRS RRF RE RDA RGA RTB g vLkw vPkw
Reference average sound level at 25 m distance under prescribed/idealized conditions Real average sound level at 25 m distance under given conditions Traffic volume Percentage of heavy vehicles in the traffic (weight > 2.8 t) Correction for the speed limit Correction for the pavement Correction for the rises and falls along the road Correction for the absorption characteristics of building surfaces Correction for the propagation absorption due to the receiver distance Correction for the attenuation due to the ground and atmospheric conditions Correction for the attenuation due to the topography and building dimensions Road gradient Speed limit for light vehicles (normally 30e130 km/h) Speed limit for heavy vehicles (normally 30e80 km/h)
[dB] [dB] [vehicles per hour] [%] [dB] [dB] [dB] [dB] [dB] [dB] [dB] [%] [km/h] [km/h]
3.2.2. Italia (CNR, 1991) The Italian CNR model (Consiglio Nazionale delle Ricerche, or National Research Council in English) was published in 1983 (Canelli et al., 1983 [121]) and then improved by Cocchi et al., in 1991 [122]. The model is a modification of the RLS (Germany) model where the traffic flow is considered as a line source. The nomenclature for this model is presented in Table 10 (Cocchi et al., 1991 [122]). The A-weighted equivalent sound level for the traffic is given by
LAeq ¼ a þ 10 log10 ðQL þ bQP Þ 10 log10
d d0
þ DLV þ DLF þ DLB þ DLS þ DLG þ DLVB
3.2.3. US (TNM, 1998) The Traffic Noise Model (TNM) is a software to predict pass-by noise levels (at a distance of 15 m) near highways. It is the only validated model approved by Federal Highway Administration (FHWA, USA) to test compliance with Code of Federal Regulations “Traffic Noise Prediction” (23 CFR 772.9(a)) (FHWA, 2012 [123]). The first version of TNM was released in 1998 (FHWA, 1998 [124]) based on extensive measurement data (over 6000 pass-by events) taken from 1993 to 1995 (Fleming et al., 1995 [125]). After several revisions, the current and mature version is TNM 2.5 (software) released in 2004. TNM Version 3.0 is under development, with 2D graphics and GIS functionality implemented (FHWA, 2013 [126]). The SPL at the receiver is adjusted from a reference sound level termed as reference energy mean emission level (REMEL) that describes the maximum sound level from a vehicle under given conditions. The possible conditions are listed in Table 11 (Rochat et al., 2012 [127]). The energy mean level (overall A-weighted) associated with the maximum level during pass-by (REMEL, for one vehicle) is given by
LA;max ðSÞ ¼ 10 log10
h i SA=10 10ðBþDEB Þ=10 þ 10ðCþDEC Þ=10
(124)
Table 10 Nomenclature for CNR model (Cocchi et al., 1991 [122]). Symbol
Meaning
Unit
LAeq QL QP d d0 DLV
A-weighted equivalent sound level for the given traffic Light vehicles in 1 h Heavy vehicles in 1 h Distance between the lane center and the observer Reference distance (25 m) Correction for the mean flux velocity (0 for 30e50 km/h; þ1 for 60 km/h; þ2 for 70 km/h; þ3 for 80 km/h; þ4 for 100 km/h) Correction for the presence of reflective facade near the observer (þ2.5) Correction for the presence of reflective facade in the opposite direction (þ1.5) Correction for the pavement type (0.5 for smooth asphalt; 0 for rough asphalt; þ1.5 for cement; þ4 for rough pavement) Correction for the road gradient Correction for the presence of traffic lights (þ1) or slow traffic (1.5) Coefficient related to noise emission from a single vehicle (default ¼ 35.1 in Italy) Coefficient related to the weightage factor for higher noise emitted by heavy vehicles (default ¼ 6 in Italy)
[dBA] [vehicles per hour] [vehicles per hour] [m] [m] [dBA]
DLF DLB DLS DLG DLVB a b
[dBA] [dBA] [dBA] [dBA] [dBA] [dBA] []
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
39
Table 11 REMEL conditions in TNM (Rochat et al., 2012 [127]). Category
Condition
Vehicle type Pavement type
Automobile; medium truck; heavy truck; bus; motorcycle Dense-graded asphalt concrete (DGAC); open-graded asphalt concrete (OGAC); Portland cement concrete (PCC); average (a combination of DGAC and PCC) Cruise; full-throttle
Operating condition
where S is the vehicle speed in mph; A, B, and C are coefficients related to the vehicle type, pavement type, and operating conditions; DEB and DEC are corrections to convert the level mean to the energy mean, and are given by
0
1 N X Li Lavg 1 C DE ¼ 10 log10 B 10 10 A @ N i¼1
(125)
where N is the number of samples, Li is the overall A-weighted level for the ith sample, and Lavg is the linear average over all samples. The REMEL for one-third octave band spectra is given by
LA;max ðS; f Þ ¼ 10 log10
h i SA=10 10ðBþDEB Þ=10 þ 10ðCþDEC Þ=10 þ ðD1 þ D2 SÞ þ ðE1 þ E2 SÞlog10 ðf Þ þ ðF1 þ F2 SÞlog10 ðf Þ2
þ ðG1 þ G2 SÞlog10 ðf Þ3 þ ðH1 þ H2 SÞlog10 ðf Þ4 þ ðI1 þ I2 SÞlog10 ðf Þ5 þ ðJ1 þ J2 SÞlog10 ðf Þ6 (126) where f is the nominal center frequency of the one-third octave band; D1, D2, …J2 are the REMEL model coefficients. The adjusted sound level at the receiver is calculated using the following equation that considers traffic flow Atraffic,i, propagation distance Ad, and shielding As.
LAeq;1h ¼ ELi þ Atraffic;i þ Ad þ As
(127)
where ELi is the vehicle noise emission for each vehicle type i depending on LA,max; Atraffic,i for each vehicle type i is given by
Atraffi;i ¼ 10 log10
Vi 13:2 Si
(128)
3.2.4. Dense-graded (Hanson and James, 2004) Hanson and James (2004) [128,129] model was developed by National Center for Asphalt Technology (NCAT) in Auburn University, sponsored by Colorado Department of Transportation (CDOT) and Federal Highway Administration (FHWA). Fortysix different pavements (each 61 m in length, 2.7 km in total) in the NCAT test track were tested, most of which were densegraded hot mix asphalt (HMA). The sound levels were measured using two approaches, close-proximity trailer (CPX) and close-proximity sound intensity (CPI). The study focused on the mixture properties of the pavements. For each pavement, seven tires were tested at a speed of 72 km/h, and the average noise level was calculated as the model output. The final regression equation for the tire noise level is
LA ¼ 93:4 2:56 AirVoids þ 0:53 FinenessModulus AirVoids
(129)
where LA is the overall A-weighted sound pressure level (OASPL), and R2 ¼ 0.64. The fineness modulus (FM) is an empirical factor obtained by adding the total percentages of a sample of the aggregate retained on each of a specified series of sieves, and dividing the sum by 100 (TXDOT, 1999 [130]; ASTM, 1980 [131]). The sieve sizes used are No. 100 (150 mm), No. 50 (300 mm), No. 30 (600 mm), No. 16 (1.18 mm), No. 8 (2.36 mm), No. 4 (4.75 mm), 3/8 in. (9.5 mm), 3/4 in. (19.0 mm), 1-1/2 in. (38.1 mm), and larger sizes, increasing in multiples of 2. The FM values of the fine aggregates are in the range of 2.00e4.00, while the course aggregates smaller than 38.1 mm have values in the range of 6.50e8.00. The pavements tested in the NCAT study have combinations of fine and coarse aggregates, and hence they have intermediate values (4.07e5.06). In addition, the separate influences of the following parameters were analyzed: fineness modulus (R2 ¼ 0.51), air voids (%) (R2 ¼ 0.11), and pavement age (R2 ¼ 0.78). The study also ranked the noise levels for four types of pavements as open-graded mixes (coarse gradations) > stone matrix asphalt matrix > dense-graded HMA > open-graded mixes (fine gradations). 3.2.5. Concrete texture (Rasmussen, 2009) Rasmussen (2009) [132] correlated TPIN with typical concrete pavement textures such as burlap drag, rake tining, and diamond grinding. 3D texture profiles were measured over 1000 sections representing approximately 400 nominal textures Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
40
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
and totaling 70 km, with RoboTex laser measurement system. Enormous amount of data (1 GB for 1.6 km) were collected owing to the high sampling frequency (1000 Hz) and high resolution (spatial resolution of approximately 0.4 mm2 and height resolution of 0.01 mm). 3D texture bridging (envelopment) filter was employed to achieve data reduction. The noise data were collected using the OBSI (On-board Sound Intensity) method (96 kHz), except that there were two additional channels for the data acquisition system (DAQ). One of the channels was used to record the event markers, i.e., signal from an optical transceiver (reflective delineator) indicating the start and end of the test section. Thus, the texture data and noise data could be synchronized after compensating for the travel time between the optical transceiver (at the road side) and the OBSI microphone (at the tire). The study conducted measurements in the frequency range of 500e5000 Hz, but focused more on 630, 800, and 1000 Hz third-octave bands. Equations were developed for the various types of concrete pavements, i.e., they have different input texture parameters and coefficients, as shown below. It was concluded that noise cannot be predicted by texture depth (Rasmussen, 2010 [133]).
8 OBSITT ¼ f Ltx;160 ; Ltx;25 ; Rk;tr > >
< OBSILT ¼ f Ltx;40 > OBSIDG ¼ f Ltx;80þ63þ50þ40 ; Skew tr > : OBSID ¼ f Ltx;50þ40 ; Skewtr ; Rk;tr
(130)
where OBSI is the overall A-weighted sound intensity level (OASIL), the subscripts TT, LT, DG, and D indicate transverse tining, longitudinal tining, diamond grinding, and drag, respectively; Ltx,XX is the texture level for third-octave band centered at a texture wavelength of XX (mm) according to ISO 13473-4 (2008) [134]; Skewtr is the texture skew in the transverse direction (unitless), according to ASME B46.1 (2009) [135]; Rk,tr is the core roughness depth in the transverse direction (mm), according to ISO 13565-2 (1996) [136]. The complete equation for the drag concrete pavement is
OBSID ¼ 91:0 þ 0:315 Ltx;40 þ Ltx;50 1:37Skewtr 3:11Rk;tr
(131)
where R2 is 0.57, and the standard error is 0.6 dBA. It may be noted that the tire used is not the current SRTT, but a Goodyear Aquatred III test tire. 3.2.6. Asperity (Fujikawa et al., 2009) Fujikawa et al. (2009) [108] developed a statistical model correlating the near-field TPIN levels with pavement roughness parameters. The roughness parameters in the final model are asperity spacing xA and asperity height unevenness hd, as illustrated in Fig. 18. Thirteen pavement profiles were measured with a single-head laser profile meter to cover 2e4 m, covering no less than 80 asperities for each sample. The specifications of the thirteen pavements are listed in Table 12. The TPIN levels were measured with a microphone placed under the vehicle beside the tire. The individual correlation between TPIN and asperity spacing was reported as R2 ¼ 0.62, with a standard error of 1.4 dB, while the individual correlation between TPIN and asperity height unevenness was reported as R2 ¼ 0.53, with a standard error of 1.5 dB. Then, multiple regression analysis was employed to correlate the TPIN levels with both the parameters. The final equation is
LR ¼ 0:25〈xA 〉 þ 1:55〈hd 〉 þ 91:8
(132)
where LR is the regressed A-weighted sound pressure level in dBA, is the average asperity spacing in mm, and is the average asperity height unevenness. The coefficient of determination was R2 ¼ 0.83 with a standard error of 0.9 dB. It was considered that the reduced spacing and height unevenness of the pavement asperities will lead to reduced TPIN. In an earlier study by the same authors, Fujikawa et al. (2004) [137], another pavement parameter, mean profile depth (MPD, in mm) was also included. The regression equation is given below.
LR ¼ 0:34〈xA 〉 þ 1:80〈hd 〉 1:15MPD þ 93:2 for block pattern tire; LR ¼ 0:33〈xA 〉 þ 1:07〈hd 〉 0:22MPD þ 91:1 for rib pattern tire:
(133)
Fig. 18. Illustration of pavement roughness parameters in Fujikawa et al. model (source from Fujikawa et al., 2009 [108], Fig. 1; reprinted under fair use provision).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
41
Table 12 Specifications of pavements measured in Fujikawa et al. model (source from Fujikawa et al., 2009 [108], Table 2; reprinted under fair use provision). No.
Pavement
Max. Age Asperity spacinsg aggregate [year] size [mm] Average 95% Confidence xA [mm] range [mm]
Asperity height unevenness
Remarks
Average 95% hd [mm] Confidence range [mm]
B0
Resin crushed stone pavement 13
3
16.7
1.2
2.73
0.28
B1
Shot blasted dense asphalt
20
3
22.2
1.8
0.74
0.09
B2 B3 B4 B5 B6 B7 B8 B9 B10
Semi-flexible pavement Semi-flexible pavement Semi-flexible pavement Dense asphalt max. 20 mm Dense asphalt max. 13 mm Dense asphalt max. 13 mm Dense asphalt max. 13 mm Dense asphalt max. 13 mm Dense asphalt max. 13 mm
13 13 13 20 13 13 13 13 13
6 6 6 3 1 5 10 1 3
17.4 20.6 20.6 17.1 10.3 13.4 11.8 11.4 11.6
1.5 2.1 2.1 1.2 1.0 1.1 0.9 1.2 1.1
0.51 0.52 0.58 0.61 0.20 0.40 0.44 0.16 0.19
0.07 0.09 0.10 0.06 0.03 0.05 0.06 0.02 0.02
B11 Dense asphalt max. 8 mm B12 Dense asphalt max. 13 mm
8 13
2 1
8.7 8.1
0.4 0.5
0.17 0.14
0.01 0.01
Sharp-edged aggregates are bound with resin. Asphalt matrix is roughened by shot blasting. Open-graded aggregates Open-graded aggregates Open-graded aggregates Aggregates with large size. Ordinary type of Japanese public road. Ordinary type of Japanese public road. Test track with reinforced asphalt. Ordinary type of Japanese public road. Rate of small aggregates is relatively higher than ordinary type. Aggregates with small size. Rate of small aggregates is relatively higher than ordinary type.
The model also analyzed tire tread vibration. However, it was measured only on three types of pavements, and no equations for the relationships were given in the study. 3.2.7. Pavement core (Reyes and Harvey, 2011) Reyes and Harvey (2011) [138] developed a regression model to predict OBSI noise levels using laboratory pavement cores. The advantage of the method is that it is easier to test a pavement core in the laboratory than to test a full-scale section in the field. The sixteen field pavement cores included those commonly used in California: five open-graded asphalt concrete (OGAC), six open-graded rubberized asphalt concrete (RAC-O), three gap-graded rubberized asphalt concrete (RAC-G), and two open-graded rubberized asphalt concrete with large aggregates (RAC-O F-mixes). The age of the pavements ranged from a few months to six years, and the air void content ranged from 9% to 22%. There were two typical diameters of the cores: 101 mm and 152 mm. Airflow resistance Rf, air void content, mean profile depth (MPD), root mean square (RMS) of the texture profile, and coefficient of uniformity Cu were investigated for their relationships with sound intensity levels. Air void content was found to have strong correlation with airflow resistance; root mean square of the texture profile was found to have strong correlation with mean profile depth; coefficient of uniformity was found to have poor correlation with sound intensity levels. The specific airflow resistance Rf (in Pa$s/m) and MPD (in m) were selected in the final model, as given below (R2 ¼ 0.52e0.86).
SIf ¼ af þ bf $MPD þ cf $log10 Rf
(134)
where SIf is the sound intensity level for the frequency band f in dBA referenced to 1012 W/m2, af, bf, and cf are coefficients, whose values are listed in Table 13 (Reyes and Harvey, 2011 [138]). The specific airflow resistance is defined the pressure difference across a core specimen divided by the airflow velocity (Rf ¼ Dp/u, in Pa$s/m). It was measured only in the region of laminar flow (0.2e10 L/min), where Rf is constant. The range of Rf Table 13 Coefficients for Reyes and Harvey model (source from Reyes and Harvey, 2011 [138], Table 1; reprinted under fair use provision). 1/3 Octave band center frequency [Hz]
af
bf
cf
500 630 800 1000 1250 1600 2000 2500 3150 4000 5000 Overall A-weighted sound intensity level
91.83 90.46 93.67 90.69 83.02 79.83 82.36 79.75 76.16 71.09 64.86 96.95
5.45 6.36 3.38 2.59 3.04 1.90 0.50 0.88 2.29 2.75 3.88 3.51
1.69 1.32 0.03 0.56 1.52 2.15 1.38 0.97 1.04 1.41 1.56 0.28
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
42
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
tested was 51e76075 Pa s/m. The MPD was obtained using a laser texture scanner (Model 9200) developed by AMES Engineering. The author has noted that only one core for each section was scanned, owing to limited funding and time available, and recommended that at least three cores should be measured according to ISO 13473-1 (1997) [139]. It is also noted that the laboratory MPD and the measured values in the field have a good correlation (R2 ¼ 0.88). It was shown that airflow resistance has a positive correlation (R2 ¼ 0.64e0.81) with sound intensity in the range of 1250e2500 Hz (high frequency). MPD has a positive correlation (R2 ¼ 0.3) with sound intensity in the range of 500e630 Hz (low frequency). However, neither of the two parameters was able to explain the noise at the 800 Hz and 1000 Hz frequency bands (R2 z 0), where most spectral content exists, and tire contributions might dominate. In addition, neither of the two parameters showed satisfying relationships with overall A-weighted sound intensity levels (R2 ¼ 0.1). 3.2.8. EU (CNOSSOS, 2012) The EU model is the proposed common European method to predict noise emission for road traffic, and was first released in 2012 (Kephalopoulos et al., 2012 [140]). The noise emission is expressed in terms of sound power level, assuming that each point source of a vehicle is 0.05 m above the surface. The model presents separate calculations for TPIN and propulsion noise, for five types of vehicles (m ¼ 1 for light vehicle, 2 for medium heavy vehicle, 3 for heavy vehicle, 4 for powered two-wheeler, and 5 for other customized vehicles). For the TPIN, the sound power level LWR,i,m for the vehicle type m and octave frequency band i (125e4000 Hz) is given by
LWR;i;m ¼ AR;i;m þ BR;i;m log10
!
v vref
þ DLWR;i;m ðvÞ
(135)
where v is the vehicle speed, vref is the reference speed of 70 km/h; AR and BR are coefficients for the respective vehicle type m and octave frequency band i; DLWR,i,m is the total correction factor considering the pavement type, studded tires, speed variation, and air temperature. For propulsion noise, the sound power level LWR,i,m for the vehicle type m and octave frequency band i is given by
LWP;i;m ¼ AP;i;m þ BP;i;m log10
v vref vref
! þ DLWP;i;m
(136)
It was shown that speed has much smaller influence on the propulsion (engine) noise than TPIN for the first three vehicle types. The overall sound power level for the vehicle is the energy sum of TPIN and propulsion noise.
0 B LW;i;m ¼ 10 log10 @10
1 LWR;i;m 10
þ 10
LWP;i;m 10
C A
(137)
The equivalent sound power level for a steady traffic flow Q (vehicle per hour) with average vehicle speed vm is given by
LW;eq;i;m ¼ LW;i;m þ 10 log10
Q 1000vm
(138)
3.2.9. Mode-wise (Syamkumar et al., 2013) Syamkumar et al. (2013) [6] developed statistical equations to predict pass-by TPIN for each combination (mode) of vehicle type and pavement. Six vehicle types were covered: heavy commercial vehicle (HCV), bus, light commercial vehicle (LCV), passenger car, auto-rickshaw (AR), and two-wheeler (2W). The pavements included eight asphalt concrete pavements (AC) and four cement concrete pavements (CC). The investigated parameters are listed in Table 14. The general form of the statistical equations is
TPINðmodeÞ ¼ a0 þ a1 p1 þ a2 p2 þ a3 p3 þ a4 p4 þ a5 m2
(139)
where TPIN is in the unit of dBA, and as an example, the coefficients for passenger cars on asphalt concrete pavement and cement concrete pavement are presented in Table 15. It was found that the TPIN levels on cement concrete pavements are 2 dB higher than those on asphalt pavements. Wet pavements increase TPIN by 3 dBA in the case of HCVs compared to dry pavements. The wind direction has influence on TPIN to the extent of 1e2 dBA in the case of HCVs. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
43
Table 14 Investigated parameters in Syamkumar's model (Syamkumar et al., 2013 [6]). No.
Parameters
Symbol
Unit
Value
1 2 3 4 5 6 7
Vehicle type Pavement type (texture depth) Loading condition Vehicle speed Wind direction Pavement temperature Pavement condition
m1 m2 p1 p2 p3 p4 p5
[] [mm] [kg] [km/h] [ ] [ C] []
HCV, bus, LCV, car, AR, 2W AC, CC Loaded (L), unloaded (UL) 10-100 km/h Towards, away (0e180 ) 20-40 C Wet, dry
Table 15 Statistical equation coefficients and R2 for passenger cars in mode-wise model (Syamkumar et al., 2013 [6]).
[TPIN(Car)]AC [TPIN(Car)]CC
a0 (Intercept)
a1 (Load)
a2 (Speed)
a3 (Wind)
a4 (Temperature)
a5 (MPD)
R2
79.10 75.43
0.0013 0.0028
0.2817 0.4343
0.0059 0.02
0.016 0.0015
0.1260 0.3137
0.918 0.911
3.2.10. FDOT (Wayson et al., 2014) The 10 FDOT model (Wayson et al., 2014 [141]) was developed by researchers at the John A. Volpe National Transportation Systems Center (Cambridge, MA), sponsored by Florida Department of Transportation (FDOT). Various types of pavements at over 45 locations in Florida (especially rigid pavements, dense-graded and open-graded pavements) were measured and ranked by the OBSI trailer (near-field) and a roadside device (far-field). The sound levels measured by the two devices have the following relationship (i.e., propagation model).
SPLWayside ¼ SILOBSI D
(140)
where SPLWayside is the road side sound pressure level in dBA, SILOBSI is the sound intensity level measured by the OBSI trailer in dBA; D is the correction term, and is given by
D ¼ 32:57 þ 0:0349$FN þ 18:094$MPD 0:0493$AG4
(141)
where FN is the average coefficient of friction for a ribbed tire, MPD is the mean profile depth, AG4 is the percent aggregate at screen No. 4 (4.74 mm). A residual standard error of 1.0 dBA was reported with R2 ¼ 0.73. For the FDOT OBSI system, the repeatability check showed an error of 0.26 dB at 95% confidence level. The pavement aging effect on TPIN was quantified as 0.2 dB/year. The frequency components of TPIN were also investigated in the study. It should be pointed out that the OBSI system was affixed to the wheel casing instead of affixing to the wheel hub, and hence the measurement might not be accurate owing to suspension compliance. Further, the direct relationship between the pavement parameters and the OBSI or pass-by levels were not reported. 3.3. Principal component analysis models Principal component analysis (PCA) uses an orthogonal transformation to identify a smaller number of uncorrelated variables, called principal components, from a larger number of possibly correlated variables. Principal component regression (PCR) is a regression analysis technique based on PCA. The advantage of PCA models is that the least number of variables (principal components) can be used to explain the maximum amount of variance. 3.3.1. Asphalt pavement (Ongel et al., 2008) Ongel et al. (2008) [142] combined the predictions for TPINs of different asphalt pavements into one universal statistical equation. OBSI tests as per the AASHTO standard (2008) [143] were conducted on 72 pavements, including open-graded, dense-graded, rubberized open-graded and rubberized gap-graded asphalt pavements. The initial parameters investigated in the study are listed in Table 16. Ten input parameters were selected based on the order of importance to develop the model. The statistics of the ten input parameters and the output parameter (OBSI) for all the measurements are listed in Table 17. The correlation matrix was generated using ordinary least-squares (OLS) multiple regression; the details are presented in Table 18. However, correlations for an input of 6 MPD were not reported. The final linear regression equation for OBSI prediction is
OBSI ¼ 107:6 þ 0:172Age 1:74Type 0:10AV 0:48IRI 1:25FM 0:005Cu þ 0:004MPD þ 1:84PR 0:09RTC þ 0:003SLT (142)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
44
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Table 16 Initial parameters investigated in Ongel model (Ongel et al., 2008 [142]). Category
Parameter
Symbol
Unit
Description
Pavement mixture
Mix type
Type
[]
Categorical variable: 1 for open-graded mixes and 0 for gapand dense-graded mixes
Age Nominal maximum aggregate size Fineness modulus
Age NMAS FM
[year] [mm] []
Coefficient of uniformity
Cu
[]
Coefficient of curvature
Cc
[]
Rubber inclusion
RI
[]
Air void content Mean profile depth Root mean square of profile deviation International roughness index British pendulum number Surface layer thickness Well gradation
AV MPD RMS IRI BPN SLT WG
[%] [mm] [mm] [m/km] [] [mm] []
Presence of transverse cracking
PTC
[]
Presence of fatigue cracking
PFC
[]
Pavement distress condition
Environmental
Presence of raveling
PR
[]
Pavement temperature
T
[ C]
On-board sound intensity level
OBSI
A measure of aggregate gradation (larger indicates coarser grade) Cu ¼ D60/D10, where D60 is the sieve size associated with 60% passing, and D10 is the sieve size associated with 10% passing Cc ¼ D30/D10, where D30 is the sieve size associated with 30% passing Categorical variable: 1 for rubberized mixes, and 0 for nonrubberized mixes A measure of macrotexture A measure of macrotexture A measure of macrotexture Categorical variable: 1 for well graded mixes, and 0 for poorly graded mixes; If Cu > 4 and 1 < Cc < 3, the mix is well graded; otherwise poorly graded Categorical variable: 1 if the total length of transverse cracks 5 m in 150 m section, and 0 otherwise Categorical variable: 1 if the total fatigue cracking 5% of the wheel path area in 150 m section, and 0 otherwise Categorical variable: 1 if the total raveled area ¼ 5% of the total area in 150 m section, and 0 otherwise Pavement temperature measured during the OBSI measurements Output parameter, corrected and reported with their equivalent values at 96 km/h
[dBA]
Table 17 Statistics of parameters in Ongel model (Ongel et al., 2008 [142]).
Input 1 Input 2 Input 3 Input 4 Input 5 Input 6 Input 7 Input 8 Input 9 Input 10 Output
Parameter
Symbol
Unit
Mean
Standard deviation
Minimum
Maximum
Mix type Age Fineness modulus Coefficient of uniformity Air void content Mean profile depth International roughness index Surface layer thickness Presence of transverse cracking Presence of raveling On-board sound intensity level
Type Age FM Cu AV MPD IRI SLT PTC PR OBSI
[] [year] [] [] [%] [mm] [m/km] [mm] [] [] [dBA]
1.5 12.4 0.04 974.6 651.4 57 5 21.1 3.8 43.8 37.01
0.7 5 0.06 328.4 196.1 7.8 0.5 14.9 2.4 22.1 15.1
0.7 4.1 0.000007 405.6 282.3 41 3.8 2.9 <0.01 10.4 8.35
3.8 21.6 0.31 1807.00 1121.60 75 5.9 58 8.7 112.5 56.4
where R2 ¼ 0.73, and the standard error is 0.93 dBA. In Table 18, it can be seen that many parameters have strong correlations with each other. Cu has significant correlations with Type, AV, and FM (absolute value > 0.7). Thus, PCR was then employed to divide these input parameters into three orthogonal (uncorrelated) groups in the order of importance. The first group is “flow properties”; it includes AV, Cu, Type, FM, and SLT (in the order of importance); the second group is “surface roughness”; it includes PR, IRI, and Age; the third group is “crack resistance”; it includes PTC and Age. The statistical equation is
OBSI ¼ 100:87 1:28 FlowProperties þ 0:79 SurfaceRoughness 0:43 CrackResistance
(143)
where R2 ¼ 0.61, and the standard error is 1.04 dBA.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
45
Table 18 Correlation matrix for nine input parameters and PBSI output in Ongel model (source from Ongel et al., 2008 [142], Table 3; reprinted with permission from American Society of Civil Engineers).
OBSI Age Type AV IRI FM Cu PR PTC SLT a b
OBSI
Age
Type
AV
IRI
FM
Cu
PR
PTC
SLT
1.000 0.284a 0.477b 0.416b 0.384b 0.347b 0.446b 0.300a 0.273a 0.281a
0.287a 1.000 0.069 0.061 0.477b 0.010 0.101 0.206 0.235 0.129
0.479b 0.069 1.000 0.767b 0.139 0.607b 0.730b 0.063 0.321b 0.555b
0.417b 0.061 0.767b 1.000 0.071 0.640b 0.763b 0.062 0.238 0.687b
0.364b 0.477b 0.139 0.071 1.000 0.018 0.113 0.410b 0.146 0.076
0.349b 0.010 0.607b 0.640b 0.018 1000 0.720b 0.049 0.306a 0.465b
0.439b 0.101 0.730b 0.763b 0.113 0.720b 1.000 0.041 0.180 0.508b
0.297a 0.206 0.063 0.062 0.410b 0.049 0.041 1.000 0.095 0.199
0.274a 0.235 0.321b 0.238 0.146 0.306a 0.180 0.095 1.000 0.165
0.283a 0.129 0.555b 0.687b 0.076 0.465b 0.508b 0.199 0.165 1.000
Correlation is significant at 0.05 level (2-tailed). Correlation is significant at 0.01 level (2-tailed).
In a further study, Ongel and Harvey (2010) [144] investigated the influence of pavement characteristics on the frequency content of tire pavement noise. The effects of pavement parameters on different 1/3 octave band frequencies are presented in Table 19. 3.3.2. MTD (Zhang et al., 2014) Zhang et al. (2014) [145,146] developed a model to predict the mean texture depth (MTD) of the pavement macrotexture based on TPIN and vehicle speed. The TPIN cannot be explicitly obtained using this model, but it provides sufficient insight into the relationship between TPIN and MTD. The noise was measured using a microphone placed underneath the vehicle behind the driver side rear tire (behind the tire, BTT method); the noise measurement included TPIN and the noise caused by wind and vehicle vibrations. The acoustic energy was obtained by integration over the frequency spectrum (DC to 2 kHz), and then PCA was applied for reduction of the effects of noise and speed (Zhang et al., 2014 [145]; Zhang et al., 2012 [147]); thus, the combined method is termed as PCA energy method. PCA is a successful method to extract features from measurements (Wang et al., 2002 [148]; Hotelling, 1933 [149]; Jolliffe, 2002 [150]). It is a statistical procedure that uses an orthogonal transformation to convert possibly correlated variables into a set of values of linearly uncorrelated variables called principal components. All the tests (using the same tire) were conducted at the National Center for Asphalt Technology (NCAT) in Lee County, Alabama (Zhang et al., 2014 [145]), as in the case of Hanson model. Employing Taylor expansion and least-squares analysis (LS), the final equation was given by
MTD ¼ 0:2 þ 1:55En þ 0:44Vn þ 0:76En Vn
(144)
where MTD is the predicted mean texture depth in mm; En and Vn are normalized PCA energy and driving speed (dimensionless), respectively, and are given by
Table 19 Effects of pavement parameters on different 1/3 octave band frequencies (Modified from Ongel and Harvey, 2010 [144], Table 3). Frequency [Hz]
500
630
800
1000
1250
1600
2000
2500
3150
4000
5000
Type Age FM Cu AB MPD a RMS IRI SLT PTC PR b Rubber c NMAS d PFC
[
[
Y
Y
Y
Y
Y [ Y
Y [ Y
Y [ Y Y Y
Y [ Y Y Y
Y [ Y Y Y
Y [ Y [ Y Y Y
Y
[ Y [ [ [ [ Y
Y [ Y [ Y Y Y
Y
[ Y [ [ [ [ Y
Y [ Y [ Y
Y [ Y Y Y
[
[ [
[ [ [
[ [
[ [
[ [
[ [
[ [
[ [
[
[ [
[ [
[ [
[
a b c d
Y
RMS is the root mean square value of profile deviation in mm. Rubber is the pavement material with rubber inclusion (coded as 1 for rubberized mixes and 0 for non-rubberized mixes). NMAS is the nominal maximum aggregate size in mm. PFC shows the presence of fatigue cracking (coded as 1 if the total fatigue cracking is 5% of the wheel path area in a 150-m section, and 0 otherwise).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
46
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
En ¼
E Emin Emax Emin
(145)
Vn ¼
V Vmin Vmax Vmin
(146)
where E is the PCA energy in Hz$Pa, V is the driving speed in m/s; the subscripts min and max indicate the minimum and the maximum values, respectively, through all the tests. It was reported that the accuracy of this model is 83.61% with good repeatability when MTD is in the range of 0.2e3 mm, which is 10% higher than the results obtained with energy method (40e400 Hz) without PCA. On the other hand, a database is always needed for PCA to perform calibration and acquire the limits of the values. 3.3.3. Thin surfacing (Li et al., 2015) Li et al. (2015) [151] developed a statistical tire/road noise model considering the texture and absorption for thin-layer surfacings. Based on different combinations of input pavement parameters, three models were developed using PCR. The three combinations are presented in Table 20. The 12 output parameters are overall noise levels and 1/3 octave band noise levels in the frequency range of 315e3150 Hz. The coefficient of determination (R2) for the three models are presented in Table 21. In general, the more the number of parameters, the more accurate the statistical model is. 3.4. Fuzzy curve fitting models A fuzzy curve fitting model is used to solve a nonlinear complex system by a few simple “If … Then” fuzzy rules (Roychowdhury, 1998 [152]). The increase in such rules gives better modeling capability, especially with the advancement of computer technology. The advantage is that no specific form of equations is needed to fit the parameters. The fuzzy logic imitates the human learning process, and has a couple of derivative applications, such as artificial neural network (ANN), neuro-fuzzy system, and fuzzy genetic arithmetic (F-GA), which will be discussed in the following sections. In this sense, the fuzzy curve fitting model can also be called neural computation model. 3.4.1. CPB ANN (Fry et al., 1999) Fry et al. developed an ANN model to correlate vehicle design parameters with vehicle-controlled pass-by (CPB) noise (Fry et al., 1999 [153]; Fry and Jennings, 2003 [154]). As per ISO 362 [155,156], the vehicle should enter the test track with an initial speed of 50 km/h, and is to be driven through with 100% acceleration. The measurements are conducted in both directions, in 2nd and 3rd gears, while the maximum sound pressure levels (Max2nd and Max3rd) are recorded for each test run. The final average drive-by noise level DBN is given by
DBN ¼
Max2nd þ Max3rd 1dBA 2
(147)
The input parameters of the ANN model cover seven functional categories: powertrain, intake, exhaust, tire, test conditions, engine bay, and vehicle performance. The binary and categorical inputs were encoded to numeric values. Seventy-five
Table 20 Input pavement parameter combinations of surface texture and sound absorption (modified from Li et al., 2015 [151], Table 3). Model
Number of input parameters
Surface texture
Sound absorption
1 2 3
15 7 10
TL250, TL125, TL63, TL32, TL16, TL8, TL4, TL2, TL1 Mean profile depth (MPD) TL250, TL125, TL63, TL32, TL16, TL8, TL4, TL2, TL1
AL1000, AL1250, AL1600, AL2000, AL2500, AL3150 AL1000, AL1250, AL1600, AL2000, AL2500, AL3150 Maximum absorption coefficient
TLx is the texture level (dB) at the wavelength of x mm (ref. ¼ 106 m). ALf is the sound absorption level at the frequency of f Hz.
Table 21 Coefficient of determination for the three models (Li et al., 2015 [151]). Model
LA,eq
L315
L400
L500
L630
L800
L1000
L1250
L1600
L2000
L2500
L3150
1 2 3
0.87 0.66 0.80
0.68 0.49 0.70
0.76 0.53 0.77
0.82 0.66 0.81
0.87 0.85 0.87
0.90 0.82 0.85
0.92 0.72 0.87
0.89 0.65 0.80
0.82 0.59 0.74
0.94 0.71 0.89
0.84 0.78 0.93
0.90 0.81 0.89
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
47
parameters were on the short list, refined from over 200 parameters by five vehicle noise experts. However, the details of these parameters were not published. The only output of the ANN model is the pass-by noise level. The suggested number of hidden neurons for an I-J-K ANN with I inputs, J hidden neurons, and K output neurons is given by (Tarassenko, 1998 [157])
J¼
pffiffiffiffiffi IK
(148)
The number of examples N should not be too small to capture the trend of the data, and also not too large to avoid overtraining or over-fitting when the ANN tries to fit the random noise. The rule of thumb is given below.
W ¼ ðI þ 1ÞJ þ ðJ þ 1ÞK N ¼ ð3 30ÞW
(149)
where W is the number of free variables, including weights and biases. The lower limit of the number of examples is given by
Nlower
pffiffiffiffiffi ¼ 3 IK I þ K þ
! rffiffiffiffi K þ1 I
(150)
Normally, K is much smaller than I; then, the equation becomes
Nlower z3I 1:5 K 0:5
(151)
It can be seen that I is more important than K. Based on the equations above, at least 2000 test runs were needed (Fry et al., 1999 [153]). To reduce this number, the ANN was divided into two submodels, as shown in Fig. 19. The vehicle acceleration was first used as the output of the vehicle performance network (20-5-1) with a mean percentage error (MPE) of ±7% (R2 ¼ 0.9) (Fry et al., 1999 [153]). Then, the output was changed to engine speed at the exit of the test track, because pass-by noise has a strong correlation with the engine speed (9 dBA per 1000 rpm). The error was found to decrease to ±4.8%. The type of neural network used is multi-layer perceptron (MLP), which is the most widely used network. The training algorithm is error back-propagation, or back-propagation. 85% of the published neural network applications has been created using some variant of the back-propagation algorithm (Wasserman, 1993 [158]). The corresponding algorithm in MATLAB is RPROP (resilient back-propagation), where the inputs were transformed to zero mean and unit variance to avoid “stuck units” (Tarassenko, 1998 [157]). Both the logsig (logistic sigmoid) and tansig (hyperbolic tangent sigmoid) transfer functions in MATLAB were tried for the hidden neurons. The best network found for the vehicle performance network and pass-by noise network is presented in Table 22. This model provides an insight into the vehicle acoustic performance in the early design stage before a prototype is built; this will save cost and engineering effort to a great extent. The prediction error of the model is within ±1.4% (i.e., ±1.1 dBA for the noise limit of 74 dBA), which is better than value that would be guessed by an experienced engineer (±3 dBA). The model also works for diesel and turbo-charged vehicles, but it is not valid for an automatic transmission vehicle, as its gear ratio will not be constant during the test. The authors have suggested a technique that combines neural networks with Bayesian statistics, called automatic relevance determination (ARD) to refine the input selection [159], and thereby potentially obtain a better prediction accuracy.
Fig. 19. Two-stage ANN models for pass-by noise (source from Fry and Jennings, 2003 [154], Fig. 3; reprinted with permission from Springer).
Table 22 Metrics for two-stage ANN models for pass-by noise (Fry and Jennings, 2003 [154]). Sub-ANN
Structure
Set for training and validation
Set for test
Accuracy
Vehicle performance network Pass-by noise network
25-15-1 58-19-1
2503 cases from 24 vehicles 4282 cases from 21 vehicles
997 cases from 13 vehicles 1424 cases from 7 vehicles
±4.8% ±1.4%
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
48
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
3.4.2. Pavement aging (Khazanovich and Izevbekhai, 2008) Khazanovich and Izevbekhai (2008) [5] presented statistical equations to predict OBSI levels for different types of pavements using the data from 51 test sections. All the tests were conducted using a standard reference test tire (SRTT) (ASTM, 2014 [160]) at 97 km/h. The input parameters were estimated single axle load (ESAL), pavement age (Age), international roughness index (IRI), and surface rating (SR). The general equations have the following form.
I ¼ ESALa Ageb IRI m SRy
(152)
where I is the predicted OBSI sound intensity level; a, b, m, and n are the model exponents. For different types of pavements, such as asphalt pavements, astroturf drag, transversely tined white topping, and turf dragged white topping, these exponents have different values, which are calculated using the universal Levenberg-Marquardt algorithm (LMA), i.e., the hybrid of steepest-descent and least-squares nonlinear model fitting techniques. Some examples are presented in Table 23. However, no global values for all pavements were presented. It was found that pavement aging increases the TPIN on asphalt pavements, but decreases the TPIN on rigid concrete pavements. It should be noted that pavement age does not have a direct influence on TPIN, but it affects IRI and SR because of the texture degradation caused by traffic loads and environmental weathering. In a later study (Izevbekhai, 2012 [161]), Izevbekhai dropped pavement age, ESAL, and SR as the input parameters, but added temperature T, asperity interval which measures the texture wavelength ASP, texture direction relative to the traffic direction DIR (categorical variable, with values 0 or 1), texture depth TD, and texture spikiness which measures the texture amplitude SP (categorical variable, with values 0 or 1). The final equation is
293 T IRI ASPT TD þC þD ðD DIR þ EÞ þ ðF DIR þ GÞ þ ½ðSPÞðJ DIR þ KÞ I ¼AþB T IRIT ASP MPD
(153)
where the capital letters A, B, C, D, E, F, G, J, K are the model constants, MPD is the mean profile depth; and IRIT and ASPT are the corresponding values at the reference temperature. The errors for the model were found to be within 1.5 dBA for 90% of the field measurements. It was also found that friction number (FN) and mean profile depth (MPD) are not significant parameters for TPIN levels. 3.4.3. F-GA (Li et al., 2009) Similar to Che et al. model (ANN), Li et al. (2009) [162] used fuzzy genetic arithmetic (F-GA) to statistically correlate the tread pattern parameters with noise levels. The input tread pattern parameters include the width and length of the tread block, the width and length of the grooves, the total number of base pitches, pitch sequencing, and row offset. The output noise level is given below.
d¼
k
X
Lp;i Lp;i;M
(154)
i¼1
where i is the index for each frequency band, Lp,i is the sound pressure level corresponding to specific tread pattern input, Lp,i,M is the sound pressure level in the noise standard line (curve M) provided by the American Test Center (ATC), and d is the model output. It was claimed that the model successfully optimized the parameters of the tread pattern for low noise tires, using fitness function iteration. However, further details on the optimized parameters and the accuracy of the model were not reported. It may be noted that both Che et al. model (ANN) and Li et al. model (F-GA) assumed that the tread pattern noise is independent of the tread block shape (only size matters). Kumar et al. (2013) [163] used multi-objective genetic arithmetic (MOGA) to optimize the tire tread pitch sequence, considering several performance parameters (noise, handling, and wear). 3.4.4. ANN (Che et al., 2012) Che et al. (2012) [164] used ANN to correlate the tire parameters with tire noise. ANN is a powerful tool to solve complex and highly nonlinear problems such as TPIN. Fig. 20 shows a schematic representation of a typical neural network. The ANN consists of multiple layers of neurons. A neuron can be mathematically interpreted as a nonlinear transfer function controlled by weights. The ANN receives the input data (tire and pavement parameters) in the “input layer”. In the
Table 23 Exponents for Khazanovich model (Khazanovich and Izevbekhai, 2008 [5]). Pavement Type
a (ESAL)
b (Age)
m (IRI)
n (SR)
Cracked white topping Bituminous (asphalt) Transversely tined
0.0075 0.188 0.00166
2.208 1.245 1.7678
0.0696 0.041 -.001858
0 0.025 0
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
49
Fig. 20. Schematic of a typical artificial neural network (ANN) (source from Che et al., 2012 [164], Fig. 2; reprinted with permission from Trans Tech Publications Ltd).
same way, the layer that outputs the results (noise spectrum in narrowband, 1/1 octave or 1/3 octave) is called the “output layer”. The input and output parameters used in this study are listed in Table 24. It may be noted that the tests were conducted using the laboratory drum (DR) method (Yong et al., 2009 [165]); the surface was considered to have no variations or runout, and hence there were no inputs of pavement parameters. The ANN statistical equation is given below.
Lp ðf Þ ¼ g Tw ; Tp ; TR ; PA ; PC ; PN ; PT ; PR ; PL ; PS ; PD ; V; L; F
(155)
where f is the center frequency of the corresponding octave band. The layers between the input layer and the output layer are normally called hidden layers. The simplest neural network is composed of an input layer and an output layer. The hidden layers improve the capacity of the network to solve complex problems. This model had one hidden layer. The number of neurons in the hidden layers does not follow any special rule, and it is chosen by the user. If the number is very small, it will be difficult for the ANN model to capture all the relationships; if the number is very large, the time for training will increase and “over-match” problem might occur. The model has the following empirical equation.
m¼
pffiffiffiffiffiffiffiffiffiffi xþlþa
(156)
where m is the number of hidden neurons, x is the number of neurons in the input layer, l is the number in output neurons, and a is a constant with value ranging from 1 to 10. The study employed back-propagation (BP) network training based on genetic algorithms (GA) that use crossover and mutation operations to determine the number of hidden neurons (6), number of training epochs (6000), and the initial weights (global optimization).
Table 24 Input and output parameters of ANN (Che et al., 2012 [164]).
Input
Parameter
Symbol
Description
Tire structure (size)
Tw Tp TR PA PC PN PT PR PL PS PD V L F Lp,125 Lp,250 Lp,500 Lp,1000 Lp,2000 Lp,4000
Tire width Aspect ratio Tire radius Area of single-pitch tread pattern block Area of single-pitch tread pattern groove Number of tread blocks per pitch Total number of pattern pitches Arrangement rule of the pitches Displacement (offset) between the pattern strips Symmetrical characteristics of tread patterns Groove depth Vehicle speed Tire load Inflation pressure Sound pressure level at the octave band of 125 Hz Sound pressure level at the octave band of 250 Hz Sound pressure level at the octave band of 500 Hz Sound pressure level at the octave band of 1000 Hz Sound pressure level at the octave band of 2000 Hz Sound pressure level at the octave band of 4000 Hz
Tread pattern
Operation
Output
Sound pressure level
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
50
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
While the network is trained, the bias and the connections (weights) between the elements are adjusted. For a specific input, the network produces a specific output. Then, the output is compared with an expected result or target. This process is repeated by adjusting the weights, until the difference between the output and the target is less than the maximum permissible error. In this study, the training data obtained from the tests were divided into two parts: training samples (50 groups) and test samples (10 groups). It was reported that the average error was 1.4%. The main advantages of the ANN over nonlinear regression analysis used in the current hybrid methods are twofold. Firstly, no knowledge of the functional relationship between the input parameters and the noise spectrum output is needed. Secondly, the amount of data needed for training is less in the ANN than in the nonlinear regression. The ANN approach can be applied to various research areas. Another study (Kostial et al., 2013 [166]) correlated the angular frequency (output) of the tires with the tire structural parameters (input), such as nominal width, profile number, rim diameter, load index, and speed index. Additional input parameters were also included, such as radial static stiffness, circumferential static stiffness, lateral static stiffness, and static torsional stiffness. It was reported that the ANN results were qualitatively comparable with FEM simulation results with a root mean square error of 4.314 Hz. Yet another study (Kongrattanaprasert et al., 2009 [167]; Kongrattanaprasert et al., 2010 [168]) correlated the road surface states (output), such as snowy, slushy, wet, and dry states, with the measured tire noise of passing vehicles (input). An accuracy of approximately 90% was reported. A similar study (Kandpal et al., 2013 [169]) correlated the vehicle categories (output) with the measured noise (input). ANN was also used for tire contact optimization to improve the handling and wearing performance (Nakajima, 1999 [170]). The input is the sidewall and tread band contour; the output is the uniformness of contact pressure distribution. El-Gindy et al (1999 [171], 2001 [172]) developed a tire/pavement contact-stress model based on ANN for truck tires. The input parameters considered were vertical load and inflation pressure, and no tire parameters were included. However, separate ANN models were developed for one radial-ply tire and one bias-ply tire. 3.4.5. Bayesian (Yu and Lu, 2013) Yu and Lu (2013) [173] developed empirical models for four asphalt pavements: dense-graded asphalt concrete (DGAC), open-graded asphalt concrete (OGAC), rubberized open-graded asphalt concrete (RAC-O), and rubberized gap-graded asphalt concrete (RAC-G). The study presented two types of models: (1) overall model for all the asphalt pavements; (2) individual model for each type of asphalt pavement. The noise data were collected over a span of four years (296 observations covering 74 sections); certain observations and circumstances were missing or not considered for unavoidable reasons. To address this issue, a multiple imputation (MI) algorithm was used to capture the unobserved heterogeneity due to the missing parameters or the lack of knowledge on the relevant parameters. A Monte Carlo Markov chain (MCMC) sampling was used to deal with the panel structured data (Lu et al., 2009 [174]; Lu et al., 2010 [175]). Bayesian simulation was adopted to allow potential parameters to be random but include all the existing knowledge on the parameters. The noise level was measured using the OBSI method (A-weighted, 500e5000 Hz). The final prediction equation for each pavement type is given by
OBSI ¼ a0 þ a1 Age þ a2 AirVoid þ a3 MPD þ a4 AADTC þ a5 Pre þ a6 Thickness þ ε
(157)
where Age is the number of years since the pavement was open to traffic, AirVoid is the air void content (%), MPD is the mean profile depth (mm, ASTM E1845 [176]), AADTC is the average daily traffic on the main lane, Pre is the annual total precipitation (mm), Thickness is the thickness of the surface mix (mm), ε is an error term; a0 to a6 are coefficients to be estimated; the results are presented in Table 25. For the overall model, the variable Thickness is replaced by NewThickness that equals the product of the thickness of the given surface mix and the corresponding probability for this mix to be chosen. It was concluded that TPIN increases with pavement age and MPD, and decreases with air void content. The TPIN levels vary in the ascending order as RAC-O < OGAC < RAC-G < DGAC. From Table 25, it can be seen that the ability to resist noise level (small a1) increase with age rank varies as RAC-O > OGAC > RAC-G. The efficiency of air-void content increase for noise absorption (large absolute value of a2) varies as DGAC > RAC-G > RAC-O > OGAC. The efficiency of mix thickness increase for noise reduction (large absolute value of a6) varies as RAC-O > OGAC > DGAC > RAC-G.
Table 25 Coefficients for Yu and Lu (Bayesian) model (Yu and Lu, 2013 [173]). Coefficient
DGAC
OGAC
RAC-O
RAC-G
Overall
a0 (Intercept) a1 (Age) a2 (AirVoid) a3 (MPD) a4 (AADTC) a5 (Pre) a6 (Thickness)
103.6 0.0137 0.28 2.29 103 3 105 4.1 105 0.029
101.2 0.177 0.12 7.11 104 6.43 104 2.17 104 0.034
102.5 0.07 0.25 1.58 103 1.55 104 1.52 104 0.02
101.1 0.187 0.25 2.55 103 4.88 104 4.2 105 0.0276
101.4 0.083 0.199 1.41 103 1.33 104 1.69 104 0.021
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
No.
Model name
Author
Year
Ref.
Method
Output
Frequency range [Hz]
Mechanism
Driver parameter
Tire parameter
Pavement parameter
Environment parameter
1
RLS (Germany)
BMVI
1990
[120]
Pass-by, Leq
N/A
Traffic
Slope
CNR (Italia)
CNR
1991
[122]
Pass-by, Leq
N/A
Traffic
N/A
Slope
3
TNM (US)
FHWA
1998
[124]
500e5000
N/A
Type
Dense-graded
2004
[128,129]
N/A
N/A
FM, voids, age
N/A
5
Concrete texture
Hanson and James Rasmussen
Vehicle and traffic TPIN
Propagation attenuation Propagation attenuation N/A
4
2009
[132]
Pass-by, 1/3 octave CPX, CPI, OASPL OBSI, 1/3 octave
Speed, vehicle type Speed, vehicle type Speed, vehicle type, acceleration 72 km/h
N/A
2
500e5000 630e1000
Texture impact
Buick Century, 97 km/h
SRTT
Texture, skewness
N/A
6
Asperity
Fujikawa et al.
2009
[108]
Nonlinear regression Nonlinear regression Nonlinear regression Nonlinear regression Linear regression, bridging filter Linear regression
BTT, OASPL, octave
500e4000
TPIN
80 km/h, 3.2 kN, 170 kPa
185/65R15, rib pattern
N/A
7
Pavement core
Reyes and Harvey
2011
[138]
Linear regression
500e5000
TPIN
N/A
SRTT
8
CNOSSOS (EU)
N/A
2012
[140]
Nonlinear regression
OBSI, 1/3 octave, OASIL Pass-by, LW, octave
Asperity spacing, asperity height unevenness, MPD Airflow resistance, MPD
125e4000
Speed, vehicle type, acceleration
Studded
Type, age
Air temperature
9
Mode-wise
Syamkumar et al.
2013
[6]
Linear regression, sensitivity analysis
CPB, OASPL
N/A
TPIN, vehicle and traffic TPIN
Speed, tire load, vehicle type
N/A
Type, wetness
10
FDOT
Wayson et al.
2014
[141]
Linear regression
OBSI, OASIL, pass-by, OASPL
500e4000
TPIN
88 km/h 97 km/h
SRTT
FN, MPD, AG4
Pavement temperature, wind direction Wind < 4 km/h
N/A
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89 51
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 26 Summary of traditional regression models on TPIN.
52
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
3.5. Summary The traditional regression models are summarized in Table 26. The regression analysis was the earliest one applied to statistical TPIN modeling. It has been widely used for traffic noise prediction by government transportation agencies. Nonlinear regression was dominant before the early 2000's, while linear regression has prevailed thereafter, probably because linear regression has easy form of equations, and needs less data to improve the accuracy. Besides pass-by noise measurement, more advanced and professional noise measurement techniques such as OBSI have been frequently used, where the 1/3 octave band noise spectrum was typically the output with a large frequency range of 500e5000 Hz. The vehicle speed was normally included in the modeling. However, unlike most deterministic models, the tire load or other tire parameters were not of much interest. Various types of pavement parameters were included, such as pavement type, slope, texture, friction, age, and wetness. Environmental parameters such as temperature and wind were also considered. However, the traditional regression model was unlikely to provide insight into the noise generation mechanisms or separate the different mechanisms. The principal component analysis models are summarized in Table 27. The input and output parameters of the model are similar to those of the traditional regression model. The principal component regression is more advanced than linear or nonlinear regression, and may include more input parameters in one model. The fuzzy curve fitting models are summarized in Table 28. Fuzzy logic technique, such as ANN, has been widely applied to various engineering problems, including TPIN modeling, since the late 1990's. The input and output parameters are similar to those of the regression models. However, the tire parameters such as tread pattern are more commonly included as input parameters. 4. Hybrid models 4.1. Introduction Hybrid models can be divided into deterministic submodels and statistical submodels. In the deterministic submodels, physical preprocessing and conditioning of input quantities are conducted to acquire intermediate parameters as inputs to the following statistical submodels. In most cases, the relationships in the statistical submodels can be significantly improved by including the preprocessing part of the data. The purpose of hybrid models is to find a compromise between the level of physical details and statistical accuracy. During recent years, the hybrid models appear to have become more popular, because they retain the accuracy of the statistical models and also provide insight into the mechanisms like the physical models.
Table 27 Summary of principal component analysis models on TPIN. No. Model name 1 2 3
Author Year
Ref.
Method
Asphalt Ongel 2008 [142] OLS, PCR pavement et al. MTD Zhang 2014 [145,146] PCA energy et al. Thin Li et al. 2015 [151] Principal surfacing component regression
Output
Frequency Mechanism Driver Tire Pavement Environment range [Hz] parameter parameter parameter parameter
OBSI, OASIL
N/A
TPIN
97 km/h
SRTT
Various
BTT, energy
0e2000
TPIN
Speed
Same tire
MTD
80 km/h
N/A
N/A Texture, MPD, absorption
CPX, 1/3 octave 315e3150 TPIN band SPL, OASPL
Pavement temperature N/A
Table 28 Summary of fuzzy curve fitting models on TPIN. No. Model name
Author
Year Ref.
1
CPB ANN
Fry et al.
1999 [153] Resilient BP
2
Pavement Khazanovich 2008 [5] Levenberg- OBSI, Aging and Izevbekhai Marquardt OASIL F-GA Li et al. 2009 [162] F-GA Octave band SPL ANN Che et al. 2012 [164] GA-BP DR, octave band SPL Bayesian Yu and Lu 2013 [173] Bayesian, OBSI, 1/3 MCMC, MI octave band, OASIL
3 4 5
Method
Output
Frequency Mechanism Driver range [Hz] parameter
CPB, OASPL
N/A
Vehicle
N/A
TPIN
125e4000 TPIN 125e4000 TPIN 500e5000 TPIN
Tire Pavement parameter parameter
Material, ISO 10844 geometry, slick included 97 km/h SRTT Type, ESAL, Age, IRI, SR N/A Tread N/A pattern Speed, tire Size, tread Standard load, inflation pattern surface Traffic N/A Age, voids, volume MPD, thickness
50 km/h (accel.), gear
Environment parameter N/A
30 C N/A N/A Precipitation
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
53
In this section, 19 hybrid models are reviewed; they consist of 9 tire-pavement interface models, 6 mechanism separation models, and 4 noise propagation models. The model name, method, output noise parameter with the noise measurement technique, noise generation mechanisms investigated, and input parameters of each model are listed and compared in tables as a summary in section 4.5. The summary tables can be used as a reference to direct the readers to the models of their interest. 4.2. Tire-pavement interface models These models calculate the excitation characteristics of the interface between the tire and pavement within the contact patch. For example, the pavement texture is converted to a roughness parameter with reference to the tire using the envelope technique. The contact force/pressure spectrum is then obtained for further analysis of the tire vibrations. The TPIN is finally correlated with the contact force/pressure spectrum (intermediate parameters) using experimental data. 4.2.1. Truck contact (Fong, 1998) Fong (1998) [177] developed a model to predict the near-field sound pressure level from the computed value of contact pressure induced by the road surface texture. On-road measurements were conducted for 175/70R13 passenger tires (smooth and patterned), as well as 7.00R15 truck tires (radial and cross-ply) at a constant speed of 45 km/h using an on-board sound pressure technique (OBSP). This model is one of the few that included truck tires. The pavements were chip seal surfaces ranging from coarse-textured Grade 2 to finetextured Grade 6 (MTD in the range of 0.85e2.63 mm). The pavement texture was measured using laser profilometer (sampling distance of 3 mm, vertical resolution of 8 mm). The sound pressure was measured around the circumference of the rear right tire at four positions, one of which is close to the center of the contact patch. The sound pressure was measured using microphones at 10 kHz with a time duration of 104 s. The contact pressure was calculated by assuming that the contact load due to the interaction with the texture asperities was equal to the tire inflation pressure. It was found that the contact pressure increased with texture levels, as was expected. Transfer functions were calculated between the output (sound levels) and input (contact pressure). It was found that the transfer functions were independent of the pavement textures, which means that the key factor that influences tire noise was actually contact pressure. A good agreement was obtained between the predicted and measured noise levels, generally within 2 dBA across all the frequencies. 4.2.2. DGAC texture (Domenichini et al., 1999) The DGAC texture model (Domenichini et al., 1999 [18]) characterized the dense-graded asphalt concrete (DGAC) pavement texture, and related it to the near-field and far-field TPINs. The porosity of the wearing course and thickness of the surface layer were kept constant or ignored in the study. The study was part of a TINO research program, the other task of which was to identify the relationship between the mix composition and the pavement texture (D'Andrea, 1999 [178]). Combining the two studies, it will be possible to control the TPIN at the pavement design phase by suitable selection of aggregate type/shape/dimensions, aggregate grading curve, binder type, etc. Eleven types of pavements were tested, including one porous pavement as a reference and two TINO prototype pavements. The texture measurement was performed according to ISO 13473-2 (2002) [179] and the Italian standard CNR BU 125/88. (1) The microtexture was measured using a British portable pendulum tester (low speed friction test); (2) the macrotexture was measured using a TINO 3D profilometer; (3) the mega-texture was measured using a TINO 3D profilometer (2e2000 cycle/m or 0.5e500 mm) and a “Dipstick” walking profilometer (0.014e3.33 cycle/m, or 0.3e70 m). Both far-field and near-field noise tests were conducted with four microphone positions. Two tires were tested (a slick tire and a normal tire) at two speeds (50 km/h and 80 km/h). The correlation between the texture at different spatial frequencies (cycle/m) and noise levels at different frequencies (Hz) is represented in the form of correlation chromatic plots as shown in Fig. 21. Absolute values between 0.75 and 1 indicate good correlations. The preliminary model used was to correlate the near-field noise with the Huschek coefficient or Huschek texture index (Huschek, 1996 [180]) (R2 ¼ 0.79). However, the Huschek coefficient did not consider pavement unevenness. As a result, a new texture index (TI) was proposed; TI is given by
TI ¼ 2:5$PSP þ LUNE
(158)
where PSP is the “peak-slope parameter”, given by
8 LPEAK $i > > < PSP ¼ 100 > lfPEAK þ lfPEAK þ lfPEAK þ > :L PEAK ¼ 3
(159)
LPEAK is the first peak of the texture level above the wavelength of 2 mm (or below the spatial frequency of 500 cycle/m), smoothed over the nearest three spectral lines, as illustrated in Fig. 22. Subscript fPEAK is the spatial frequency corresponding Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
54
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Fig. 21. Example of pavement texture-rolling noise correlation chromatic plot (source from Domenichini et al., 1999 [18], Fig. 10; reprinted under fair use provision).
Fig. 22. Texture index (TI) illustration for Domenichini et al. model (source from Domenichini et al., 1999 [18], Fig. 14; reprinted under fair use provision).
to wavelength lPEAK where the peak occurs. Subscripts fPEAK and fPEAKþ are the previous and following spatial frequencies, respectively. lxx indicates the texture level at the spatial frequency of xx cycle/m. The letter i refers to the slope of the descending portion above fPEAK in the macrotexture region, and is given by
i¼
LPEAK L500 log10 ðlPEAKÞ log10 ð2Þ
(160)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
55
where L500 is the texture level at the spatial frequency of 500 cycle/m (wavelength of 2 mm) smoothed over the nearest three spectral lines. LUNE is the unevenness level, and is given by
LUNE ¼
l1:26 þ l1:58 þ l2:00 3
(161)
LUNE represents the average texture level centered on 630 mm (500e800 mm). The final regression equations for the far-field sound pressure level at 50 km/h and 80 km/h are
SPL50km=h ¼ 0:47$TI þ 42:94 SPL80km=h ¼ 0:47$TI þ 49:39
(162)
where R2 ¼ 0.91. 4.2.3. SPERoN (Beckenbauer and Kuijpers, 2001) Beckenbauer and Kuijpers (2001) [181] developed a statistical and physical model to predict coast-by (CB) noise levels based on pavement parameters. The model is called SPERoN, which is the short form for statistical physical explanation of rolling noise. The noise data were based on the “Sperenberg project” (Influence of the road surface texture on the tire/road noise) for Federal Highway Research Institute of Germany (BASt), which included coast-by spectra of 840 tire-pavement combinations (21 dense surfaces; 12 normal, 2 slick, and 2 grooved passenger car tires) at speeds in the range of 50e120 km/h. The coast-by noise was measured 7.5 m away from the center of the vehicle, 1.2 m above the ground. The model was developed on a stepwise basis; hence, one additional input variable was included to expand the model for each step, and multivariate regression analysis was employed. The following requirements have to be met for each step: (1) the variables are independent; (2) the addition of a new variable must increase the coefficient of determination R2; (3) the influence of the variables on the noise spectra must be explainable physically. The final equation of the model includes four input variables (R2 ¼ 0.8): speed, frequency-transformed contact pressure spectrum, tire width, and shape factor; the equation is given below.
LpAF;i ¼ cvAF;i $10 log10
v þ ccAF;i $LcF;i þ cwAF;i $w þ cgAF;i $Lg v0
(163)
where the subscript i is the index for the third octave band, LpAF,i is the A-weighted sound pressure level in dB, v is the vehicle speed in km/h, v0 is the reference vehicle speed (80 km/h), LcF,i is level of the frequency-transformed contact pressure in dB, w is the tire width in mm, Lg is the shape factor level in dB, cvAF,i is the regression coefficient for the speed term, ccAF,i is the regression coefficient for the contact pressure term, cwAF,i is the regression coefficient for the tire width term, and cgAF,i is the regression coefficient for the shape level term. The model is hybrid, because input parameters LcF,i and Lg are not available at the beginning, but deterministically calculated based on the other tire-pavement parameters, as shown in Fig. 23.
Fig. 23. Scheme of SPERoN model (source from Beckenbauer and Kuijpers, 2001 [181], Fig. 1; reprinted under fair use provision).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
56
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
The shape factor g is derived from the Abbot curve, and is used to identify convex and concave (or positive and negative) textures. As illustrated in Fig. 24(a), a shape factor of 25% in convex shape means that if an imaginary horizontal plane is lowered from the top of the surface by a distance R/2, it will be in contact with 25% (PCL, percentage contact length) of the texture. The linear scale shape factor g has the following relationship with the log scale shape level Lg (dB).
Lg ¼ 20 log10
g 1%
(164)
The contact pressure distribution was calculated based on the algorithm developed by Clapp et al. (1988) [182]. A contact patch length of 100 mm was assumed for a stationary tire. The contact pressure pc(x) as a function of the longitudinal position x was derived from the pavement texture profile and the average tire tread radial stiffness s. Assuming the tire rolling speed is v, the spatial domain of the contact pressure was transformed to time domain.
pc ðtÞ ¼ pc
x v
(165)
Then, the frequency domain of the contact pressure pc(f) was obtained using Fourier transformation of pc(t). The linear scale pc(f) has the following relationship with the log scale level of contact pressure LcF(f) in dB.
LcF ðf Þ ¼ 20 log10
pc ðf Þ pc0
(166)
where pc0 ¼ 106 N/m2. It was concluded that the spectral energy decreases in the range of 800e1250 Hz as the speed increases, especially for rough pavement, as shown in Fig. 25(b). The coefficient of determination R2, the regression coefficient for the shape level term cgAF, and the regression coefficient for the speed term cvAF are shown in Fig. 26. It can be seen in Fig. 26(a) that slick tires have the highest coefficient of determination (R2 z 0.9) for most of the frequencies, while for the patterned tires, the values are relatively low, especially in the range of 400e800 Hz (R2 z 0.7), where the tread impact is assumed to have more effect. It is also concluded that the introduction of shape level Lg improves the reliability (R2 increases by 0.04) especially at low frequencies (<500 Hz), as shown in Fig. 26(a). Beckenbauer and Kuijpers (2001) [181] also concluded that the shape level Lg is independent of other input variables such as the level of the frequency transformed contact pressure LcF, because different values of Lg do not change the contact pressure regression coefficient ccAF. In Fig. 26(c), cvAF,i can be considered as the speed exponent that indicates the type of noise and corresponding generation mechanism. For slick tires, the speed exponent is greater than 4, indicating air pumping as the dominant mechanism. For patterned tires, the speed exponent is much lower, especially in the range of 400e800 Hz (cvAF z 3), again indicating the tread impact mechanism. In summary, the first version of SPERoN has the best predictability for slick tires, relatively good for groove tires, but only fair for normally profiled tires. Thus, the model is valid only on dense pavements with isotropic textures for slick tires. 4.2.4. Enhanced SPERoN (Kuijpers and Blokland, 2003) Kuijpers and Blokland (2003) [183] improved the SPERoN model by including the effects of tire tread profiles. The original goal of the SPERoN model was to evaluate acoustic performance of pavements, and to develop low-noise pavements, and not low-noise tires. After the incorporation of tread profile parameters, the model became more universal and unified.
Fig. 24. Illustration of shape factor (a. asphalt with surface dressing; b. hot rolled asphalt) (source from Beckenbauer and Kuijpers, 2001 [181], Fig. 2; reprinted under fair use provision).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
57
Fig. 25. Contact pressure level spectra for different speeds (a. stone mastic asphalt 0/8; b. surface dressing 3/5) (source from Beckenbauer and Kuijpers, 2001 [181], Fig. 3; reprinted under fair use provision).
Fig. 26. (a) Coefficient of determination R2, (b) Regression coefficient for the shape level cgAF, and (c) Regression coefficient for the speed term cvAF depending on third-octave canter frequency fT. Parameter: type of tire. The dashed line in (a) is valid for the statistical model without the variable Lg (shape level) for patterned tires (Source from Beckenbauer and Kuijpers, 2001 [181], Fig. 4; reprinted under fair use provision).
The contact model in the enhanced SPERoN model is shown in Fig. 27. The tread profile was divided into parallel slices across the tire width. The static contact pressure distribution pc(x, y) for each slice at the lateral position of y was calculated using the corresponding tread profile and pavement texture profile with the same algorithm used by Clapp et al. (1988) [182]. The texture profile (both 2D texture profile and 3D tread profile) was assumed to be identical for all slices. Then, the contact pressure in the frequency domain for each slice pc(f, y) was calculated using Fourier transformation. The input parameter pc(f) for the later statistical regression model was the average of pc(f, y) for all slices. It was concluded that the introduction of tread profile parameters reduced the prediction error by approximately 0.25 dBA in the frequency range of 500e1000 Hz. The tests were conducted on a very smooth grinded concrete surface with a specific profiled tire, where the tire profile was assumed to have more importance. However, the model prediction ability decreased at other frequencies (125e400 Hz and 1000e1500 Hz). In the case of tires with different tread profiles (slick, grooved, and normal), the overall prediction ability improved (the prediction error decreased by approximately 0.5 dBA) over the entire frequency range above 400 Hz. It can be stated that as a unified model for all types of passenger car tires, the enhanced SPERoN is promising.
Fig. 27. 3D contact model in the enhanced SPERoN model (source from Kuijpers and Blokland, 2003 [183], Fig. 1; reprinted under fair use provision).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
58
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
The enhanced SPERoN also includes mechanical impedance of the pavement and tire rubber hardness as input parameters for the statistical regression model. In the same study, the static contact model was compared with the dynamical contact model by Kropp et al. (2001) [184], where the variation in contact stiffness was substituted for the tire profile effect. To be more specific, if the contact area reduces because of changes in the tread profile, the deformation amplitude increases, i.e., the contact stiffness decreases. It was found that the static contact model and dynamical contact model gave comparable results, and a speed exponent of 3 in the dominant frequency range was predicted for both models. However, it may be noted that Kuijpers and Blokland (2003) [183] also concluded that tread profiles do not have influence on the rating of acoustic qualities of pavements. Caution should be exercised before using this enhanced SPERoN model, because tread profile analysis is very complicated and time consuming. As another enhancement for SPERoN, Blokland and The (2007) [185] included more types of pavements, other than dense surfaces. Several further development attempts were performed by Chalmers and Müller-BBM within the German project “Quiet Traffic”, the EC sixth framework project ITARI and the DEUFRAKO project P2RN [186]. The information on the new advanced model (version 3.0) can be found in Table 29. This advanced model includes four TPIN mechanisms, i.e., vibration due to impact (p2vibr), airflow due to air pumping (p2airflow), cavity due to cavity resonance (p2cavity), and aerodynamic noise due to air turbulence (p2aerodyn). The mean square value of coast-by sound pressure p2coast-by is given by
8 > p2coastby ¼ p2vibr þ p2airflow þ p2cavity þ p2aerodyn > > > > > p2 ¼ aFc2 Ga1 Ba2 Sa3 > > < vibr p2airflow ¼ b Fc2 G1:5 S2 > > > g1 2 > > > pcavity ¼ cGpattern > > d1 : p2 aerodyn ¼ dV
b1
Ba2 V 4
(167)
where a, b, c, and d are the regression coefficients for the corresponding 3rd octave band, a1, a2, a3, b1, g1, and d1 are exponents; Fc is the contact force in N, G is the contact air flow resistance in Pa$s/m, B is the tire width in m, S is the tire tread stiffness in N/m, Gpattern is the spectral power for the tread pattern variation, and V is the vehicle speed in m/s. The most recent version of SPERoN has also included parameters of acoustic absorption and mechanical impedance. 4.2.5. HyRoNE (Klein and Hamet, 2007) Very similar to SPERoN, the HyRoNE (hybrid rolling noise estimation) model developed by Klein and Hamet (2007) [187] is also a hybrid model to predict pass-by noise levels as operational tools with less computational effort. The work was performed by INRETS (Hamet and Klein, 2004 [188]) and LCPC in the French PREDIT project “Texture&Bruit” and in the DEUFRAKO project P2RN (Beckenbauer et al., 2008 [186]). Both HyRoNE and SPERoN models physically analyze the raw texture of the road surface and convert it to the relevant roughness parameter with reference to the tire. The difference is that the former uses shape factor, while the latter uses the Table 29 Comparison of SPERoN and HyRoNE models (Beckenbauer et al., 2008 [186]; Klein et al., 2008 [190]).
Vehicle and tire Pavement parameter
SPERoN
HyRoNE
Passenger car tires (34 types), truck tires (Kuijpers et al., 2007 [191]) Airflow resistance, shape factor
One vehicle/tire combination (Renault Scenic 2.0 with Michelin Energy XH1/E3A tires) Acoustic absorption, porosity, shape factor and airflow resistance, porous layer thickness Twelve independent profiles (evenly distributed on both sides of the noise measurement point and characterize a 12-m long section, 6 profiles on each wheel track spaced by 1 m) Envelopment One specific tire (Michelin Energy XH1/E3A)
Texture profile
Quasi 3D texture (more than six 2D texture profiles of the pavement in parallel, 2 m each, covering 100 mm width in total, 1 mm resolution)
Texture analysis Tire parameter
Shape factor Tire type, load, tension, bending stiffness in circumferential and cross directions, tread stiffness (shore hardness) and tread pattern, radial uncoupled contact forces, inflation pressure Impulse response functions Average 3rd octave band spectrum of the contact forces As input parameter, arbitrary between 30 and 120 km/h Dense pavementsa Coast-by, 1/3 octave Maximum 2 dBA error
Tire model Contact model Speed Limitation Output Accuracy Computing time Software application Application a
210 s (69e120 km/h); 330 s (50e68 km/h) Windows (version 3.0), results in xml file Design of existing or artificial surfaces, at laboratory or test track scale
One specific tire (Michelin Energy XH1/E3A) Envelopment before spectral analysis Three nominal speeds: 70, 90, 110 km/h Isotropic or quasi-isotropic surface Controlled pass-by, 1/3 octave ±2.4 dBA, ±2.0 dBA, and ±2.1 dBA at 70, 90, and 110 km/ h, respectively 25 s (110 km/h); 45 s (90 km/h); 65 s (70 km/h) Windows (version 1.0), results in xls file Procurement process and in-situ monitoring at road scale
The most recent version of SPERoN has included acoustic impedance, and hence it is also applicable to porous pavements.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
59
method of envelopment before spectral analysis. The envelopment procedure is based on a static contact model between the pavement texture and an elastic half-space representing the tire rubber [189], as shown in Fig. 28. The tire rubber is characterized by its Young's modulus E. In the texture profile processing, wavelengths longer than 1 m and the upward peaks are removed. Another important difference between the two models is that HyRoNE also addresses attenuation mechanism (pavement absorption). HyRoNE mainly includes two TPIN mechanisms, tire vibration (100e1250 Hz) and air pumping (1250e5000 Hz), which are clearly separated in the physical and statistical submodels. In general, SPERoN is more complicated and universal than HyRoNE. A comparison of SPERoN and HyRoNE models is presented in Table 29 (Beckenbauer et al., 2008 [186]; Klein et al., 2008 [190]). 4.2.6. Road excitation (Rustighi et al., 2008) Rustighi et al. (2008) [192] developed a three-dimensional element model to predict the stochastic tire vibration due to tire/road excitation and the interior and exterior noises. In this model, it is assumed that a smooth soft tire is travelling on a road that is not very rough. The interaction between tire and the ground is modeled as a linear Winkler bedding (Johnson, 1985 [193]). The force f on the finite element points within the contact patch, as illustrated in yellow circles in Fig. 29, is given by
f ¼ Kðd wÞ
(168)
where d is the vertical displacement of the road, w is the vertical displacement of the tire, and K is the linear contact stiffness matrix. The nodal displacement vector v is governed by the following equation.
Fig. 28. Illustration of envelopment procedure in HyRoNE model (Beckenbauer et al., 2008 [186], Fig. 5; reprinted with permission from Mr. Philippe Herzog on behalf of SFA, France).
Fig. 29. Mesh used for the tire model. The circles indicate the chosen contact points (Source from Rustighi et al., 2008 [192], Fig. 2; reprinted with permission from Elsevier).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
60
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
" K2
# v2 v 2 þ K þ K j u G u M vðf; uÞ ¼ fðf; uÞ 1 0 vf vf2
(169)
where f is the tire circumferential angular coordinate, u is the angular frequency, K2, K1, and K0 are the elastic force matrices, M is the inertial force matrix, and G is the fluid-structure coupling matrix. In the frequency domain, the vertical displacement d of the road is expressed as an excitation spectral density Sd(f), given by
Sd ðf Þ ¼
c f 2:5 v v
(170)
where f is the excitation frequency in Hz, c is a constant, and v is the vehicle speed. A two-dimensional Gaussian model is used to represent the road surface, of which the roughness is homogeneous and isotropic (Robson, 1979 [194]). This approximation is experimentally determined, which is the reason why this model falls under the category of hybrid models. The calculated rms velocities of the tire wall and tread surface are shown in Fig. 30. The calculated sound pressure level (SPL) is shown in Fig. 31. The calculated interior sound pressure levels at eight hearing points inside the car cavity are shown in Fig. 32. The peaks around 60 Hz and 100 Hz may correspond to the first two internal air cavity resonances of the car, while the significant peak at approximately 230 Hz could be related to the first tire cavity resonance mode. 4.2.7. Viscoelastic contact (Dubois et al., 2011) The viscoelastic contact model [195] investigated the frictionless viscoelastic contact between a viscoelastic half-space (tire) and a rigid rough multi-indenter surface (pavement) using a macroscale approach (Cesbron et al., 2011 [196]). The viscoelastic behavior of the tire has to be considered owing to the time-dependent behavior of rubber, and the large contact area that includes many road asperities in different positions during rolling. The viscoelastic contact law for the kth asperity was described in terms of the convolution integral according to Radok's technique, as given below.
Zt Pk ðtÞ ¼
jðt tÞ
d ½f ðd ðtÞÞdt dt k k
(171)
0
where Pk is the contact force on the summit of the kth asperity, j is the relaxation function, fk the elastic force-displacement law of the kth asperity, and dk is the local approaching displacement. Then, the TPIN levels were statistically correlated with the contact forces calculated above. The linear regression equation is
Fig. 30. Root mean squared velocities of the tire elements at 80 km/h. Note that the distances ±0.1 m in the radial direction correspond to the tread section of the tire, and the distances from ±0.1 to ±0.2 m correspond to the sidewalls (Source from Rustighi et al., 2008 [192], Fig. 7; reprinted with permission from Elsevier).
Fig. 31. Calculated value of sound pressure level (SPL) at the surface of the tire elements at 80 km/h (source from Rustighi et al., 2008 [192], Fig. 9; reprinted with permission from Elsevier).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
61
Fig. 32. Sound pressure level at eight passengers' ears at 20 (continuous line), 40 (dotted line), and 80 (dashed line) km/h (source from Rustighi et al., 2008 [192], Fig. 13; reprinted with permission from Elsevier).
LN ðfto Þ ¼ aj LF ðfto Þ þ bj
(172)
where LN is the TPIN level, LF is the contact force in the frequency domain, and fto is the center frequency of each third-octave band; aj and bj are the regression parameters. The TPIN levels and contact forces are positively correlated in the frequency range of 315e1000 Hz (R2 ¼ 0.93, better than the correlation between TPIN levels and pavement texture). It was found that the viscoelastic contact forces are greater than those obtained using the elastic contact method, and the contact duration is shorter, which fits better with the experimental results and gives better prediction for tire noise. A previous study (Cesbron et al., 2009 [197]) correlated the pavement texture spectrum, contact force spectrum, and the noise level spectrum in 1/3 octave band for a slick tire on six different pavements, as shown in Fig. 33. It was found that at 30 km/h, the levels of the resultant contact forces around 800 Hz are positively correlated with the texture wavelengths in the range of 4e100 mm (macrotexture) and with the noise levels in the range of 500e1000 Hz. It was also noted that the contact length under rolling condition is 20% smaller than that under the static condition, which is because of the viscoelasticity effect of the tire rubber. 4.2.8. Contact distribution (Dubois et al., 2013) Dubois et al. (2013) [198] improved the viscoelastic contact model by using two-scale iterative method (TIM) to obtain the contact pressure distribution. The tire/road noise measurements were conducted with close proximity (CPX) method on ten surfaces of the reference test track of IFSTTAR (French institute of science and technology for transport, spatial planning, development and networks) in Nantes, France. Two tires were tested: (1) patterned tire (Michelin Energy E3A 195/60R15); (2) slick tire (Michelin Racer Slick 186/57R15). The speed effect was expressed using the following regression equation.
Lm* N
fj ¼ aj log10
V Vref
! þ bj
(173)
where Lm* N is the reprocessed measured noise level at any speed for frequency fj, V is the speed, and Vref is the reference speed m (90 km/h). An example of the data set for coefficients aj, bj, standard deviation sm j , and correlation coefficient rj is presented in Table 30. The 3D pavement texture profiles were used to calculate the contact pressure distribution. This process consists of two steps: (1) calculate the contact forces at the tips of the asperities (macroscale) (Dubois et al., 2012 [199]); (2) calculate the contact pressure within the tire-road interface (microscale). To obtain the spectral contact force, the pressure distribution p(x, y, t) in the time domain is integrated to get the total contact force F(y, t) in the time domain. dy yþ Z 2
dx
Z2
Fðy; tÞ ¼
pðx; h; tÞdxdh ydy 2
(174)
dx 2
The narrow band contact force spectrum SF(f) is given by Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
62
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Fig. 33. Iso-correlation curves between (a) contact force and noise level; (b) pavement texture and contact force; (c) pavement texture and noise level (source from Cesbron et al., 2009 [197], Fig. 17e19; reprinted with permission from Elsevier). Dy
1 SF ðf Þ ¼ Dy
Z2
SF ðy; f Þdy
(175)
D2y
where Dy is the scaling coefficient. Then, the third octave contact force level in dB is given by
LeF ðfi Þ ¼ 20 log10
SF ðfi Þ Fref
! (176)
where fi is the center frequency of the third octave band, and Fref is the reference force (1 107 N). The relationship between the noise level and contact force is given by
LeN ðfi Þ ¼ ai LeF ðfi Þ þ bi
(177)
It was found that the relationship is valid (R2 > 0.64) only at low frequencies (315e1250 for a slick tire, and 315e800 Hz for a patterned tire). The coefficients ai, bi, standard deviation sei , and correlation coefficient rei are presented in Table 31 for a slick tire and in Table 32 for a patterned tire. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
63
Table 30 Parameters of linear regression for speed effect on surface DAC 0/10 (source from Dubois et al., 2013 [198], Table 1; reprinted with permission from Elsevier). fj [Hz]
315
400
500
630
800
1000
aj bj [dB] sm j [dB] rm j
43.3 84.6 ±1.57 0.88
39.0 83.8 ±1.40 0.88
30.2 84.9 ±1.26 0.84
19.5 87.0 ±1.24 0.72
18.3 90.8 ±1.1 0.74
35.0 93.7 ±1.1 0.9
Table 31 Parameters for the correlation between noise level and contact force for slick tire (error ε ¼ 1.50%) (source from Dubois et al., 2013 [198], Table 2; reprinted with permission from Elsevier). fi [Hz]
315
400
500
630
800
1000
1250
ai bi [dB] sei [dB]
1.59 39.6 ±1.8 0.94
1.38 25.2 ±1.5 0.94
1.68 50.1 ±1.8 0.92
1.73 52.8 ±1.3 0.97
1.88 62.8 ±1.3 0.97
2.10 80.1 ±2.3 0.95
1.86 61.3 ±3.5 0.83
rei
Table 32 Parameters for the correlation between noise level and contact force for patterned tire (error ε ¼ 1:34%) (source from Dubois et al., 2013 [198], Table 3; reprinted with permission from Elsevier). fi [Hz]
315
400
500
630
800
ai bi [dB] sei [dB]
1.47 33.3 ±2.0 0.83
1.51 38.1 ±1.6 0.90
1.86 65.9 ±1.7 0.90
1.84 61.8 ±1.3 0.94
1.38 21.5 ±0.9 0.96
rei
It was also concluded that the estimation of noise levels is better by using contact force as input parameter than by using pavement texture. 4.2.9. 3D envelopment (Klein and Cesbron, 2016) The 3D envelopment model (Klein and Cesbron, 2016 [200]) is an improvement over HyRoNe model (Klein and Hamet, 2007 [187]), where the 2D enveloped texture profiles were used to calculate the tire-pavement contact forces without considering the lateral/transverse texture inhomogeneities. It was found that the 2D model did not work well for smooth pavements such as the ISO surface. To address this issue, in Ref. [200], both the 3D pavement texture and tread pattern profile were considered, and a new 3D envelopment procedure was developed for the prediction of the tire belt radiation noise. The input for the 3D envelopment calculation is the 3D combined tire-pavement roughness, obtained by subtracting the tire tread height from the road texture height, as shown in Fig. 34. The tire-pavement contact is modeled as an elastic half-space (tire) between a rigid body (pavement). The deflection of an elastic body w is given by the Boussinesq's formula.
wðx; yÞ ¼ ∬ pðu; vÞgðx; y; u; vÞdudv C
gðx; y; u; vÞ ¼
1 n2 1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pE 2 ðx uÞ þ ðy vÞ2
(178)
Fig. 34. Top left: tire tread pattern; top right: pavement roughness; bottom: combined tire-pavement roughness (longitudinal/lateral spatial resolution ¼ 1 mm, source from Klein and Cesbron, 2016 [200], Fig. 10; reprinted under fair use provision).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
64
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
where p(x,y) is the pressure distribution within the contact patch C, E is the Young's modulus of the tread rubber, and n is Poisson's ratio. The global equilibrium condition is given by
∬ pðx; yÞdxdy ¼ F
(179)
C
where F is the total tire load. The iterative algorithm can be used to solve the contact patch C, the contact pressure p, and the half-space surface displacement w. An example of the calculated contact pressure is shown in Fig. 35. The surface displacement w can be considered as the 3D enveloped combined tire-pavement roughness, which can then be spectrally correlated with the tire noise. Good correlations were observed in the frequency range of 250e1250 Hz. However, the accuracy of the model at higher frequencies (over 1250 Hz), where the air pumping mechanism is dominant, needs to be improved. The error in the 1/3 octave noise level is within 3 dB, while the error in the overall noise level is within 1 dB. The results for one patterned tire (Michelin Energy E3A 195/60R15) tested on 11 non-absorptive pavements were presented, suggesting that more tires of different tread patterns need to be validated. Some other studies on the modeling of pavement envelopment can be found in Pinnington, 2012 [201]; Pinnington, 2013 [202]; and Pinnington, 2016 [203]. 4.3. Mechanism separation models In these models, an equation form for each noise generation mechanism is assumed based on its physics. Then, the total TPIN is obtained by summing up the values from all the generation mechanisms. Finally, the parameters in the assumed equations are statistically determined via experimental data. 4.3.1. Tire/road coupling (Cao et al., 2008) Cao et al. (2008) [204] studied the coupling noise during tire-road interaction in both time domain and frequency domain. The prediction model for tire tread pattern noise assumed a patterned tire running on a smooth road. Five generation mechanisms were included, which are discussed below. (1) Air pumping at the leading and trailing edges of the contact patch: The sound pressure due to this mechanism is given by
P1 ðtÞ ¼ gb $Ab $sinðub t þ qÞ$jB$sin ux tj
(180)
where
( gb ¼
1 sinðub t þ qÞ 0
x
sinðub t þ qÞ < 0
x2ð0; 1Þ
1 1 B2 Ab ; Ab 5 9
ux 2ð8ub ; 10ub Þ
(2) Air pumping in the tread grooves:
Fig. 35. Contact pressure distribution (red curve: patterned E3A tire; black curve: slick tire) (source from Klein and Cesbron, 2016 [200], Fig. 13; reprinted under fair use provision). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
P2 ðtÞ ¼ ðPmax $sin aÞ$gs $sinðus t þ qÞ
65
(181)
where a is the groove angle, Pmax is the maximum sound pressure, and
1 sinðus t þ qÞ 0 h sinðus t þ qÞ < 0 h2ð0; 1Þ gs ¼
(3) Pipe resonance:
P3 ðtÞ ¼
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffi sin ut d$l$A$ZðdÞ $sinð10uÞ $ d$l$A$ZðdÞ$sinf½0:8u±0:2RðtÞ$tg þ jsin utj 10
(182)
where d is the groove width, l is the groove length, Z(d) is a function of d, and R(t) is a random number between 1 and 1.
8 < 1; A¼ 1 : ; 3
sin ut 0 sin ut < 0
(183)
(4) The sound pressure due to cavities and pores in the contact patch is P4(t). (5) The sound pressure due to air turbulence is P5(t). The prediction model for road noise assumed a smooth tire running on a normal road. Four mechanisms were included. (6) Smooth tire on a smooth road (base noise):
P6 ðtÞ ¼
R ðtÞ R ðtÞ $A1 $sin u1 0:5 þ 2 $t 0:5 þ 1 2 2
(184)
(7) Stick/slip (friction):
P7 ðtÞ ¼
R ðtÞ A2 R ðtÞ $ $t $sin u2 0:5 þ 4 0:5 þ 3 2 2 x,k
(185)
(8) Road unevenness:
P8 ðtÞ ¼
R ðtÞ R ðtÞ $ðA3 $h$jÞ $sin u3 0:5 þ 6 $t 0:5 þ 5 2 2
(186)
(9) Texture impact:
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
66
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
P9 ðtÞ ¼
R ðtÞ R ðtÞ $ðA4 $h$gÞ $sin u4 0:5 þ 8 $t 0:5 þ 7 2 2
(187)
Where the terms R1(t) to R8(t) are random numbers between 1 and 1, the terms A1 to A4 are the sound pressure amplitudes, and the terms u1 to u4 are the sound pressure frequencies. Then, total sound pressure in time domain is given by
PTR ðtÞ ¼ z1 P1 ðtÞ þ z2 P2 ðtÞ þ z3 P3 ðtÞ þ z4 P4 ðtÞ þ z5 P5 ðtÞ þ z6 P6 ðtÞ þ z7 P7 ðtÞ þ z8 P8 ðtÞ þ z9 P9 ðtÞ ¼ ½ z1 / z9 ½ P1 / P9 T
(188)
where z1 to z9 are the best-fit coefficients determined by experiments. Then, the total sound pressure in frequency domain HAll(f) after accounting for horn effect and pavement absorption is given by
HAll ðf Þ ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 aðf Þ$Kðf Þ$HTR ðf Þ
(189)
where a(f) is the pavement absorption coefficient, K(f) is the horn effect coefficient, and HTR(f) is the Fourier transform of PTR(t). However, the experimental data and model accuracy have not been reported. 4.3.2. Mechanism decomposition (Dare, 2012) Dare (2012) [13] decomposed the OBSI spectrum into three constituent spectra, representing the contributions from sidewall vibration, tread band vibration, and tangential vibration, as illustrated in Fig. 36. It is assumed that the shapes of the constituent spectra will not change significantly, though the magnitudes may change. The general constituent spectra were derived using SRTT on six different pavements at five speeds (16e42 km/h). For the sidewall vibration constituent, the gain function GIV,sidewall for the relationship between the measured sound intensity and vibration level is given by
GIV;sidewall ðf Þ ¼
Isidewall ðf Þ jVsidewall ðf Þj2
(190)
where f is the frequency, Isidewall is the sound intensity for sidewall vibration constituent measured with tread band covered in sound-absorbing insulation, and Vsidewall is the velocity spectrum measured with an accelerometer mounted on the sidewall. Similarly, the gain function for the tread band vibration is given by
Fig. 36. Constituent spectra and predicted total spectrum compared to measured spectrum for cell 1, leading probe. Blue: tread band spectrum. Red: sidewall spectrum. Green: tangential spectrum. Dotted black: predicted total spectrum. Solid black: measured spectrum (Source from Dare, 2012 [13], Figure 5.10; reprinted with permission from Dr. Tyler Dare of Pennsylvania State University, USA). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
GIV;treadband ðf Þ ¼
67
Itreadband ðf Þ
(191)
jVtreadband ðf Þj2
where Itreadband is measured with the sidewall covered with sound-absorbing insulation, and Vtreadband is measured with an accelerometer mounted on the tread groove. The tangential vibration is the result of stick-slip and tangential impact. The tread block was modeled as a dipole source, with two out-of-phase monopole sources at the front and rear edges of the tread block. The sound pressure P generated by a dipole is given by Ref. [97].
1 Q P ¼ r0 c 2 l
! 1 ejud þ r rþd
(192)
where r0 is the air density, c is the speed of sound, l is the wavelength of sound, r is the distance between one monopole source and receiver (2 cm), and d is the distance between the two monopole sources (length of the tread block). Q is the source strength, given by Q ¼ whV, where w is the tread block width (2.54 cm), h is the tread block height (0.64 cm), and V is the tangential velocity of the tread block. The gain function is
Gdipole
! 1 wh 1 ejud þ ¼ r0 c 2 l r rþd
(193)
The sound intensity equations for the three constituent spectra at the leading edge are given by
8 2 > > < Isidewall;lead ¼ GIV;sidewall Vsidewall;lead 2 Isidewall;trail ¼ GIV;sidewall Vsidewall;trail > > :I tangential;lead ¼ Gdipole Vtangential;lead Gtrain Ghorn;close þ Gdipole Vtangential;trail Gtrain Ghorn;far
(194)
where Gtrain is the calculated impulse train spectrum due to the frequency of tread block impact, Ghorn,close and Ghorn,far are the measured horn effect gain functions for close and far OBSI probe locations, respectively. The sound intensity levels for the three constituent spectra at the leading edge are given by
! 8 Isidewall;lead > > > ¼ 10log þ Dsidewall;lead L > sidewall;lead 10 > Iref > > > > > ! > < Itreadband;lead Ltreadband;lead ¼ 10log10 þ Dtreadband;lead > Iref > > > > ! > > > Itangential;lead > > > þ Dtangential;lead : Ltangential;lead ¼ 10log10 Iref
(195)
where Iref is the reference sound intensity equal to 1 pW/m2, and D is the prediction offset/correction for a specific pavement. The total sound intensity level at the leading edge is
0 Ltreadband;lead B Lsidewall;lead 10 10 Ltotal;lead ¼ 10 log10 B þ 10 þ2 @10
1
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Lsidewall;lead
10
10
Ltreadband;lead
$10
10
cos f þ 10
Ltangential;lead 10
C C A
(196)
where f is the measured phase difference between the sidewall and tread band sources. The average error for one-third octave band between the predicted and measured values is 2.3 dB for the leading probe, and 2.8 dB for the trailing probe. The sound levels at the trailing probe are calculated in a similar manner. To predict the total sound intensity for different pavements, a statistical model was developed.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
68
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
8 > > > Ltreadband þb1 xþb2 > > > 10 Itreadband ¼ 10 > > > < Lsidewall þb3 xþb4 10 Isidewall ¼ 10 > > > Ltangential þb5 xþb6 > > > 10 > Itangential ¼ 10 > > :
(197)
where b1 to b6 are best-fit coefficients, and x is a pavement parameter such as MPD, texture level at a given wavelength, or absorption. It was found that MPD, texture level at 12.5 mm wavelength, texture level at 3.15 mm wavelength, and skewness are the best predictors for this one-parameter OBSI model. The statistical model can also include two parameters simultaneously, as given below.
8 > > > Ltreadband þb1 x1 þb2 x2 þb3 > > > 10 Itreadband ¼ 10 > > > < Lsidewall þb4 x1 þb5 x2 þb6 10 Isidewall ¼ 10 > > > Ltangential þb7 x1 þb8 x2 þb9 > > > 10 > I ¼ 10 > > tangential :
(198)
It was found that texture levels at 12.5 mm and 3.15 mm wavelengths are the best two predictors, with an error within 0.9 dB at the leading probe and 1.1 dB at the trailing probe for the overall sound intensity levels. It may be noted that the experimental data did not cover tests at high speeds (the speeds were lower than 42 km/h). Extrapolation with speed exponent was used to account for the effect at high speeds. 4.3.3. Air pumping separation (Winroth et al., 2013) Winroth et al. (2013) [205] separated the noise component due to tire vibrations and the component due to air pumping. It was pointed out that the sound generation due to tire vibrations is proportional to the square of vehicle speed, while the sound generation due to air pumping is proportional to the fourth power of vehicle speed. Therefore, the sound pressure can be expressed as
p2rms ðf ; VÞ ¼ A2 ðf Þ$V 2 þ A4 ðf Þ$V 4
(199)
where prms is the rms value of sound pressure, f is the frequency, and V is the vehicle speed; A2 and A4 are coefficients as a function of frequency; A2 and A4 are statically determined by curve fitting from the spectra of controlled pass-by (CPB) levels in the speed range of 50e120 km/h. Several pavements were investigated: mastic asphalt, stone mastic asphalt sealed with synthetic resin, and a sand paper surface. However, only slick tire was considered in this model. It was found that air pumping noise is dominant at frequencies over 1000 Hz and at some low frequencies around 400 Hz. 4.3.4. Deterministic SEA (Mohamed and Wang, 2016) Mohamed and Wang (2016) [206] investigated tire cavity resonance using a combination of deterministic analysis (impedance compact mobility matrix method, ICMM) and statistical energy analysis (SEA). The deterministic part focuses on the cavity pressure response and the compliant wall vibration velocity response at low frequencies, while the statistical part focuses on the responses at high frequencies. In the mid-frequency range, the two methods are combined to identify the characteristics of tire-cavity coupling. The authors claimed that the same analysis method could be applied to other similar structural-acoustic systems with toroidal shape. The tire-cavity coupling system is modeled as an annular cylinder, where the side and inner walls are assumed to be rigid, while the outer shell (referred to as tire structure hereafter) is assumed to be flexible. The vibration energy E1 of the tire structure (large modal density) is statically calculated, while the acoustic energy E2 of the tire cavity (small modal density) is deterministically predicted.
E1 ¼
P1 þ uE2 hh12 ððuuÞÞh12 uðh1 þ h12 Þ
(200)
where P1 is the power input to the tire structure, u is the angular frequency, h1 and h2 are the damping loss factors of the tire structure and tire cavity, respectively; h12 is the coupling loss factor from the tire structure to the tire cavity. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
E2 ¼
1 1 ∭ V V2
Zu2 2 pWðR R Þ2 2 o i dudV pu ðr; q; zÞ r0 c2
69
(201)
u1
where V is the volume of the cavity, u1 and u2 are the lower and upper limits of the frequency band, respectively; pu is the time- and space-averaged rms pressure of the cavity obtained from experiment; r, q, and z are the cylindrical coordinates; W is the tire width, Ro and Ri are the outer and inner radii of the tire cavity, respectively; r0 is the air density, and c is the speed of sound in air. It can be seen that the coupling in this model is a strong coupling, where the structure and the fluid influence each other (the equation for E1 contains E2). 4.3.5. Statistical vibration (Le Bot et al., 2017) The statistical vibration model (Le Bot et al., 2017 [207]) was based on the deterministic model of a prestressed orthotropic plate on visco-elastic foundation (Kropp, 1999 [53]), as illustrated in Fig. 37. Both modal analysis and statistical analysis were performed to investigate the tire vibration characteristics at high frequencies. The plate is of length a and width b, unwrapped in the tire circumferential and transverse directions, respectively. The governing equation for out-of-plane vibrations is
2 4
qffiffiffiffiffi v2 pffiffiffiffiffi v2 Bx þ By vx2 vy2
!2
v2 v2 Tx 2 þ Ty 2 vx vy
!
3 v v2 5 v ¼ f ðx; y; tÞ þSþc þm vt vt
(202)
where v(x, y, t) is the displacement field (output response) in the direction normal to the plate, and f(x, y, t) is the contact force field (input excitation). In Eq. (202), m is the mass per unit area in kg/m2; Tx and Ty are the x-component and y-component, respectively, of the tensile force (in N/m) imposed by the air pressure (always positive); Bx and By are the bending stiffnesses (in N$m) in the x-direction and y-direction, respectively; S is the stiffness per unit area (in N/m3) of the elastic foundation, and c the viscous damping coefficient (in N$s/m3). The periodic boundary conditions are applied for the x-direction, while simply supported boundary conditions are applied for the y-direction. The mode shapes are obtained as
x y sin jp ; a b x y jij2 ðx; yÞ ¼ bij sin 2ip sin jp ; a b qffiffiffiffiffiffiffiffiffiffiffi 8 < aij ¼ 2=ab; i¼0 . p ffiffiffiffiffi ffi : aij ¼ bij ¼ 2 ab; is0
jij1 ðx; yÞ ¼ aij cos 2ip
i ¼ 0; 1; / j ¼ 1; 2; / i ¼ 1; 2; / j ¼ 1; 2; /
(203)
The eigenvalues are given by
mu2ij
" #2 " 2 # pffiffiffiffiffi 2ip2 qffiffiffiffiffi jp2 2ip 2 jp ¼ Bx þ By þ Tx þ Ty þS a b a b
(204)
Fig. 37. Tyre-road model. Left: the tyre is virtually unwrapped. Right: equivalent prestressed orthotropic plate under visco-elastic foundation (Source from Le Bot et al., 2017 [207], Fig. 1; reprinted with permission from Elsevier).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
70
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
The transfer mobility, i.e., frequency response function between a force applied at x0(x0, y0) and the vibrational velocity at x(x, y), is given by
x x0 y y sin jp sin jp 0 cos 2i p X 4εi b a b Yðx; x0 ; uÞ ¼ iu ab m u2ij þ ihij uuij u2 i0;j1 ( 1=2; i ¼ 0 εi ¼ 1; i>0
(205)
The identification of the physical parameters can be performed using a simplex search method from the point mobility measurement results. The dispersion relationship is given by
kðuÞ ¼
ffi vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi u ffi u 2 2 uT þ T 4B S mu t
2B qffiffiffiffiffi 2 pffiffiffiffiffi BðqÞ ¼ Bx cos2 q þ By sin2 q
(206)
TðqÞ ¼ Tx cos2 q þ Ty sin2 q where k is the wavenumber, and q is the wave propagation angle with respect to x-direction. The cut-on frequency is given by
ucuton ¼
rffiffiffiffiffi S m
(207)
The vibration characteristics are defined by three statistical parameters: modal density n(u) that indicates the size of the population of resonant modes, modal overlap M(u) that measures the overlapping of successive modes, and normalized attenuation factor m(u) that indicates the number of reflections undergone by a ray during its time life. These parameters are given by
nðuÞ ¼
2m
Z Z
Re½Yðx; x; uÞdudAx
pDu A Du
MðuÞ ¼ hðuÞunðuÞ 3 2 Z2p 1 hu exp l dq5 mðuÞ ¼ ln4 2p cg
(208)
0
where A is the plate area ab, Du is the bandwidth of interest (e.g., octave band centered at 1 kHz, i.e., [707, 1414] Hz), Y(x, x; u) is the point mobility at x, and Re denotes the real part; h(u) is the mean damping loss factor at frequency u; cg is the group speed, defined as du/dk. Another statistical parameter that characterizes the uniformity of spatial repartition of energy is s, which is defined as the ratio of spatial standard deviation of local energy and mean energy; the equation for s is given below.
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi u1 Z u s¼ Eðx; u0 Þ2 dAx Eðu0 Þ2 t Eðu0 Þ A 1
1 Eðu0 Þ ¼ A
Z
A
Eðx; u0 ÞdAx
A
Eðx; u0 Þ ¼ S0
Z
(209)
mjYðx; x0 ; uÞj2 du
Du
where S0 is the power spectral density of the force f at x0. Based on these statistical parameters, the vibrations in tires can be classified into three regimes: modal, diffuse field, and free field regimes, as shown in Figs. 38 and 39. For real tires, the damping is relatively high, and increases with frequency from 0.05 to 0.3. Therefore, as the frequency increases, the modal regime directly transitions into the free field regime without reaching the diffuse field regime. It may be noted that at low frequencies below 400 Hz, the tire dynamics is dominated by a Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
71
Fig. 38. Relative standard deviation of spatial energy repartition in the frequency, damping loss factor plane under point-excitation. Left side: modal regime with a few modes contributing to the dynamics. Top right-hand corner (black): free field regime where rays are highly absorbed. Bottom right-hand corner (bright): diffuse field regime. The broken lines are regime boundaries determined by N(u) ¼ n(u)$Du and m(u) (Source from Le Bot et al., 2017 [207], Fig. 7; reprinted with permission from Elsevier).
Fig. 39. Examples of energy fields corresponding to crosses in Fig. 38. (a) Regime of modal field (point A); (b) Regime of diffuse field (point B); (c) Regime of free field (point C) (Source from Le Bot et al., 2017 [207], Fig. 8; reprinted with permission from Elsevier).
few modes (typically six nodes), whose modal overlap is relatively low. At high frequencies, e.g., 1000 Hz, the dynamics is mainly driven by bending effects, where the modal density and modal overlap are so large (e.g., mode count N ¼ 192 for the octave at 1 kHz) that the behavior of individual modes is not distinguishable on point mobility. To investigate the non-modal domain, the geometrical acoustics approach is used instead of conventional modal analysis. The energy field may be described in terms of rays that propagate in straight lines with a speed depending on the direction. 4.3.6. Network resonator (Wang and Duhamel, 2017) In this work (Wang and Duhamel, 2017 [208]), the grooves between the tire tread and pavement texture were modeled as acoustic network resonators, as shown in Fig. 40. A numerical method for the calculation of resonant frequencies of network resonators was proposed. First, the end corrections for the open-ended pipes were calculated. The end correction dsq for the transverse pipe, assuming rectangular flat flanges as shown in Fig. 40, is given by Dalmont's method (Dalmont et al., 2001 [209]).
dsq ¼ dsq∞ þ
" 5 # asq
a a dsq0 dsq∞ þ 0:057 sq 1 sq aeff bsq bsq bsq
dsq∞ ¼ 0:811aeff dsq0 ¼ 0:597aeff
(210)
2asq aeff ¼ pffiffiffi
p
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
72
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Fig. 40. A network (in red) between a round surface and a rigid plane surface (ABCD indicates the contact patch; 1 is a longitudinal pipe while 2 and 3 are transverse pipes; bsq is half the flange width for pipe 2) (Source from Wang and Duhamel, 2017 [208], Fig. 2; reprinted with permission from Elsevier). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
where 2asq (¼ 0.009 m) is the pipe width, and 2bsq is the flange width. The end correction d for the longitudinal pipe, assuming a cylindrical flange, is obtained by BEM, and is statistically given by
d ¼ 0:00808 þ 0:22128w 3:72112w2 þ 19:80897w3
(211)
where 2w is the flange width. It may be noted that this d is slightly dependent on frequency, and Eq. (211) gives the mean values in the frequency range of 0e2000 Hz. Next, the resonant frequencies of the network resonators were solved using finite element software Abaqus, where the end corrections were accounted for during physical modeling. Finally, the genetic algorithm (GA) was used to optimize the acoustic network (e.g., pipe positions and junction types) to obtain the targeted resonant frequencies, and the result was validated by experiments on wooden network resonators. It may be noted that the acoustic resonator was used in this work to attenuate the sound pressure level, which is different from the common understanding that pipe resonance amplifies the tire noise. 4.4. Noise propagation models In these models, the sound power of TPIN is acquired statistically. Sound propagation theory is then applied to obtain the sound pressure level (SPL) at a given location. 4.4.1. UK (CRTN, 1988) The UK model (CRTN, Calculation of road traffic noise) was issued by the UK Department of Transport (DOT) in 1988 [210], and is widely used for traffic noise calculation in United Kingdom, Ireland, Australia, New Zealand, and Hong Kong (Leung and Mak, 2008 [211]). A webpage calculator is also available (NPL, 2005 [212]). Like NMPB 2008, CRTN also includes emission and propagation models. The road is modeled in CRTN as a line source instead of a set of point sources (in NMPB); additional methods to deal with dual source lines, median barriers, and corrections for thin surfaces were presented in 2008 (UK Department for Transport, 2008 [213]). The sound level is expressed in terms of L10 (the noise level exceeded for 10% of the measurement period) for the 18 h period between 06:00 and 24:00 (LA,10,18h) without noise assessment during night between 24:00 and 6:00. The source emission level is expressed as follows.
L10;18h ¼ 29:1 þ 10lgðQ Þ þ CorrectionV&p þ CorrectionG þ CorrectionTD
(212)
where Q is the number of vehicles for the 18 h duration, including all vehicles (light and heavy); CorrectionV&p is correction for mean traffic speed V (km/h) and the percentage of heavy vehicles p (%); CorrectionG is the correction for the road gradient G (%), CorrectionTD is the correction for the texture depth (TD) in mm. The equations for the three corrections are
500 5p þ 10lg 1 þ 68:8 CorrectionV&p ¼ 33lg V þ 40 þ V V
(213)
CorrectionG ¼ 0:3G
(214)
CorrectionTD;concrete ¼ 10lgð90TD þ 30Þ 20 CorrectionTD;asphalt ¼ 10lgð20TD þ 60Þ 20
(215)
CorrectionTD is 1 dBA for speeds less than 75 km/h. LA,10,18h must be converted to Lden and Lnight indicators for noise mapping assessment (Abbott and Nelson, 2002 [214]). However, the noise data at night were not collected in the CRTN method. Even though the noise prevails at night, the correlation between L10,night and Leq,night is not good enough owing to insufficient traffic volume. Another disadvantage of CRTN Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
73
compared to NMPB 2008 is that it does not consider different traffic flow types, and the fact that speeds for light vehicles and heavy vehicles are normally different. In different locations, CRTN needs to be recalibrated, which is not an easy task. For example, Table 33 shows the several recalibrated CRTN models proposed by local researchers in Hong Kong (Leung and Mak, 2008 [211]). They all have certain errors. Based on the CRTN results, Abbott and Nelson (2002) [218] correlated Leq and L10 using regression method. O'Malley et al. (2009) [219] then adapted it to an Irish scenario [36], and expressed Lden as
Lden ¼ 0:86 LA;10;18h þ 9:86dB
(216)
Murphy and King (2011) [220] noted that on an average, it underpredicts noise levels by 0.01 dB with an rms. error of 1.6 dB for the case in Dublin, Ireland. 4.4.2. France (NMPB, 1996) NMPB (1996) [221] model was described in French standard XPS 31-133, and was recommended by Environmental Noise Directive (END) for traffic noise calculation. The first version of the model is “NMPB-Routes-96 (SETRA-CERTU-LCPCCSTB)” (1996) [222], which described only the sound propagation model, while the sound emission model was presented in “Guide du Bruit” (1980) [223]. The revised model NMPB 2008 [224] separates the TPIN and engine noise (Dutilleux, 2013 [225]), and tra, 2009 [226]). is currently used in France; both noise emission and noise propagation models are used (Se In NMPB 2008, each road is modeled as a line source, or several lines in the case of multiple lanes. Each line is then divided into uninform sections; each section is divided into a set of incoherent point sources. For the noise emission model, the sound emission level E in dB for a single vehicle is given by
E ¼ LW 10 log V 50
(217)
where LW is the A-weighted sound power per meter in a given octave band in dB, and V is the vehicle speed. LW depends on the traffic conditions, as presented in Table 34. The alternative equation for sound emission level is
E ¼ E0 þ a log10
V V0
(218)
where the coefficients E0 and a are determined by the combination of traffic conditions; V0 is the reference speed (20 km/h). The sound power level per meter of lane for each elementary period k is given by
LAW=m;k ¼ Elv;k þ 10lg Qlv;k þ Ehv;k þ 10lg Qhv;k
(219)
where Q is the number of vehicles (traffic load); the subscripts lv and hv refer to light vehicle and heavy vehicle, respectively. After considering the pavement types, the average sound power level per meter of lane is given by
LAW=m;J ¼ 10lg
X
10
LAW=m;k þJ
k
10
$
hk hp
(220)
Table 33 Recalibrated CRTN models proposed by local researchers in Hong Kong (Leung and Mak, 2008 [211]). Reference
Author
Year
Equation
[210] [215] [216] [217]
CRTN Lam and Tam To et al. Tang et al.
1988 1998 2002 2004
LA,10,18h ¼ 10log10(Q) þ 33log10(V þ 40 þ 500/V) þ 10log10(1 þ 5p/V) þ 0.3G e 26.6 LA,10,18h ¼ 1.0534 LA,10,18h,CRTN 5.3347 LA,10,18h ¼ 10.6log10(Q) þ 7.1log10(pV) e 17log10(V) þ53 LA,10,18h ¼ 10log10(Q) þ 42.6log10(V þ 40 þ 500/V) þ 10log10(1 þ 5p/V) þ 0.46G e 49.2
Table 34 Traffic conditions in NMPB 2008 [224]. Conditions
Category
Vehicle type Speed Traffic flow type Road slope
Light < 3.5 t; heavy > 3.5 t <20 km/h set to 20 km/h; 20e120 km/h Fluid continuous; pulsed continuous; pulsed accelerating; pulsed decelerating Rising > 2% upward; falling > 2% downward; level 2%
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
74
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
where J is the correction for pavement type, as presented in Table 35; hk is the length of the elementary period in hours and hp is the length of the reference period in hours. The sound power level of an elementary source i in a given octave band j (125e4000 Hz) is given by
LAWi ¼ LAW=m;J þ 20 þ 10 log10 ðli Þ þ Rj
(221)
where li is the length of the elementary source line (uniform section), and Rj is the spectral correction for each octave band [227], as listed in Table 36. For the noise propagation model, geometric divergence, atmospheric absorption (ISO 9613-1, 1993 [228]), ground absorption, diffraction, and reflection are considered, assuming that each elementary source is 0.5 m above the ground. The noise propagation is divided into two distinct types: meteorological conditions favorable to sound propagation, with subscript F; and neutral (homogeneous) meteorological conditions, with subscript H. The sound levels for the two conditions are given below.
LF ¼ LW Adiv Aatm Agrd;F Adif ;F LH ¼ LW Adiv Aatm Agrd;H Adif ;H
(222)
The sound level is often expressed as Lden, which is appropriate for noise mapping for environment assessment (Murphy and King, 2014 [118] [229]). 4.4.3. Highway (Dai and Lee, 2010) Dai and Lee model (2010) [230] is called highway prediction model (HPM), and is used to predict highway traffic levels on pavements of asphalt concrete (AC) and asphalt rubber concrete (ARC). This model was developed based on ray acoustics approach that is also applied in ISO 9613 (attenuation of sound during propagation outdoors, with accuracy up to ±3 dBA within a range of 1000 m) (ISO 9613-1, 1993 [228]; ISO 9613-2, 1996 [231]). The advantage of this model is that it is presented in the form of Microsoft® Office Excel® 2007, and the users are required to provide only the parameter inputs that are very easy to acquire, such as weather conditions, pavement types, pavement porosity and permeability, hourly traffic volume, vehicle nominal speed, and receiver's distance from the road. Three assumptions were applied in the model: (1) The road is assumed to be an infinite line source; (2) Free field condition is assumed for the source and the receiver; (3) There is no barrier between the source and the receiver. The sound pressure level Lft(DW) at a given distance is given by
Lft ðDWÞ ¼ LW þ DC A
(223)
where the nomenclature is as presented in Table 37; the total attenuation is given by
A ¼ Adiv þ Aatm þ Agr þ Abar þ Amisc
(224)
4.4.3.1. Attenuation due to geometric divergence. The attenuation due to geometric divergence is calculated using the following equation.
Adiv ¼ 10 log10
Q 4pd2
(225)
Table 35 Correction for pavement type in NMPB 2008 [224]. Pavement
Noise level correction J [dB]
Porous surface Smooth asphalt (concrete or mastic) Cement concrete and corrugated asphalt Paving stones
1 for 0e60 km/h; 2 for 61e80 km/h; 3 for 81e130 km/h 0 þ2 þ3
Table 36 Values for spectral correction of each octave band [227]. j
1
2
3
4
5
6
Octave band center frequency [Hz] Rj
125 14.5
250 10.2
500 7.2
1000 3.9
2000 6.4
4000 11.4
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
75
Table 37 Nomenclature for highway prediction model (HPM) (Dai and Lee, 2010 [230]). Symbol
Unit
Description
Lft(DW) LW DC A Adiv Aatm Agr Abar Amisc Q d
[dB] [dB] [dB] [dB] [dB] [dB] [dB] [dB] [dB] [m2] [m] [dB/m] [Hz] [K] [K] [Hz] [Hz] [kPa] [kPa] [] [] [kPa] [K] [dB] [dB] [dB] [] [] [] [] [dB] [vehicles/h] [%]
Sound pressure level at a receiver location for a given octave band center frequency f and temperature T Sound power level Directivity correction (set to 0 for free field condition) Octave-band attenuation due to multiple effects Attenuation due to geometric divergence Attenuation due to atmospheric absorption Attenuation due to ground absorption Attenuation due to barriers (set to 0 because of the assumption that there are no barriers) Attenuation due to miscellaneous effects Noise source directivity index (set to 2 for hemispherical noise source) Distance from source to receiver Atmospheric attenuation coefficient Central frequency for a specific octave band Ambient temperature Reference air temperature (293.15 K) Oxygen relaxation frequency Nitrogen relaxation frequency Atmospheric pressure Reference atmospheric pressure (101.325 kPa) Molar concentration of water vapor Ambient relative humidity Saturation vapor pressure Triple-point isotherm temperature (T0 þ 0.01 K) Ground attenuation in the source region Ground attenuation in the receiver region Ground attenuation in the middle region Ground absorption coefficient Pavement porosity Pavement permeability Material coefficients determined by measurements Attenuation due to traffic volume Traffic volume Percentage of trucks in the traffic volume
a f T T0 frO frN pa pr h hr psat T01 As Ar Am
ap fp kp A1, A2, A3 Atra Vtra Ptruck
4.4.3.2. Attenuation due to atmospheric absorption. The attenuation due to atmospheric absorption is calculated using the following equation.
ad 1000
Aatm ¼
(226)
where the atmospheric attenuation coefficient a depends on weather parameters (temperature and relative humidity), and is given by
"
a ¼ 8:686f
2
11
1:84 10
pa pr
1 0:5 2:5 #h 1 1 i 2239:1 33352 T T f2 f2 0:01275e T frO þ þ þ 0:1068e T frN þ T0 T0 frO frN (227)
where the oxygen relaxation frequency fro and nitrogen relaxation frequency frN are given below.
frO ¼
pa pr
frN ¼
0:5 T 0:02 þ h 24 þ 4:04 104 h T0 0:391 þ h " 1 # 1 0
0:5 4:17 B pa T B9 þ 280he @ T0 pr
(228)
3
T T0
1
C C A
(229)
The molar concentration of water vapor in the above equations is given by
h ¼ hr
psat pr
pa pr
(230)
where the saturation vapor pressure is calculated using the following equation. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
76
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
1:261 psat ¼ 10 pr
6:8346
T T01
þ4:6151
(231)
4.4.3.3. Attenuation due to ground absorption. The attenuation due to ground absorption is given by
Agr ¼ As þ Am þ Ar
(232)
The specific calculations vary for different central frequencies of the octave bands f. However, except at 31.5 Hz and 63 Hz, the ground absorption coefficient is important. It is related to pavement parameters and given by
ap ¼ A1 ln fp þ A2 ln kp A3
(233)
4.4.3.4. Attenuation due to the barrier. Abar is set to 0, because it is assumed that there are no barriers between the source and the receiver. 4.4.3.5. Attenuation due to miscellaneous effects. The attenuation due to traffic volume is considered for two pavement types: asphalt concrete (AC) and asphalt rubber concrete (ARC). The equations are 2 2 Atra;AC ¼ 0:0001Vtra 0:0446Vtra 0:0026Ptruck þ 0:2284Ptruck þ 25:934
(234)
2 2 Atra;ARC ¼ 0:00009Vtra 0:0359Vtra 0:0011Ptruck þ 0:1132Ptruck þ 10:1336
(235)
4.4.3.6. Results. The model (HPM) was compared with the measured noise using statistical pass-by method (SPB), computeraided noise abatement (CadnaA) simulation software (CadnaA), and traffic noise model (TNM®) “LookUp” program developed by the Federal Highway Administration (FHWA), United States. The results are shown in Fig. 41 (Dai and Lee, 2010 [230]). Comparing HPM and SPB, larger errors (up to 4 dBA) occur at 7:00 a.m. and 8:00 p.m. It is probably due to drastic changes in ambient temperature, leading to atmospheric inversions (Sheadel et al., 2007 [232]). On an average, HPM is the most accurate model (having error within 1.65 dB to 1 dB) among the three prediction models. The TNM accounts for only densegraded asphalt concrete (DGAC, a type of AC), and is applicable only when the distance between the source and receiver is in the range of 10e300 m. The CadnaA always underpredicts the noise levels, which is possibly because it is based on “Road traffic noise pollution guidelines” (RLS-90, published by German Federal Ministry of Transportation), and not applicable in local areas (Canada).
Fig. 41. Comparisons between HPM and SPB, CadnaA, and TNM (source from Dai and Lee, 2010 [230], Fig. 1; reprinted with permission from Ms. Beth Darchi on behalf of ASME).
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
No. Model name
Author
Year
Output
Frequency Mechanism range [Hz]
Driver parameter Tire parameter
Pavement parameter
Environment parameter
1
Truck contact
Fong
1998 [177] Transfer function
OBSP, 1/3 octave band
10e4000
TPIN
45 km/h
Passenger and truck
N/A
OBSP, pass-by, OASPL CB, LpA,max, 1/3 octave band SPL
22e5500
TPIN
Speed
Slick & normal
Beckenbauer and Kuijpers
Cross-correlation, regression 2001 [181] Multivariate regression analysis
Texture, contact pressure spectrum Texture index
2
DGAC texture
Domenichini et al. 1999 [18]
3
SPERoN
125e5000 TPIN
50-120 km/h
Slick, tire width
4
Enhanced SPERoN
Kuijpers and Blokland
2003 [183] Multivariate regression analysis
CB, LpA,max, 1/3 octave band SPL
125e5000 TPIN
50-120 km/h
5
HyRoNE
Klein and Hamet
2007 [187] Envelopment, regression
CPB, LpA,max, 1/3 octave band SPL
100e5000 TPIN
70, 90, 110 km/h
6
Road excitation
Rustighi et al.
2008 [192] FEM, BEM, 2D Gaussian
Vibration, SPL
10e400
7
Viscoelastic contact Dubois et al.
8 9
Ref.
Method
2011 [195] Macroscale, multiasperity Contact distribution Dubois et al. 2013 [198] Weighted linear regression 3D envelopment Klein and Cesbron 2016 [200] Envelopment, linear regression
Near field, 1/3 octave band SPL CPX, 1/3 octave band CPX, 1/3 octave band; CB, Lpmax
20-100 km/h Texture impact, radiation, interior noise 315e1000 Texture impact 90 km/h 315e1250 Impact
50-110 km/h
250e1250 Belt vibration
65-110 km/h
Shape factor, contact pressure spectrum Tread pattern, tire Shape factor, contact pressure width, rubber spectrum, hardness mechanical impedance Michelin Energy Enveloped XH1/E3A texture, absorption, airflow resistance Smooth Texture excitation Slick 186/57R15 Slick and Patterned Tread pattern
Texture, contact force spectrum Texture, contact force spectrum Texture
N/A N/A
20 C
20 C
N/A
N/A
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
N/A N/A
77
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 38 Summary of tire-pavement interface models on TPIN.
78
No.
Model Name
Author
Year
Ref.
Method
Output
Frequency range [Hz]
Mechanism
Driver parameter
Tire parameter
Pavement parameter
Environment parameter
1
Tire/road coupling Mechanism decomposition
Cao et al.
2008
[204]
N/A
SPL, narrowband
0e5000
Various
N/A
Tread pattern
Texture, absorption
N/A
Dare
2012
[13]
Nonlinear least squares
OBSI, 1/3 octave band
500e5000
Speed
SRTT
Texture levels, skewness, absorption, friction
N/A
Winroth et al. Mohamed and Wang Le Bot et al.
2013
[205]
Curve fitting
125e4000
50-120 km/h
Slick
Texture
N/A
2016
[206]
ICMM, SEA
N/A
N/A
[207]
0e4000
Slick 155/70R13
Point source
N/A
2017
[208]
Modal analysis BEM, FEM
Stationary, shaker 29 psi
205/65R15
2017
CPB, 1/3 octave band SPL SPL, energy, narrowband Mobility
Sidewall, tread band, tangential vibration Tire vibration, air pumping Cavity resonance Vibration
Lab, SPL
200e2000
Pipe resonance
Stationary
Tread pattern
Smooth
N/A
2
3 4 5 6
Air pumping separation Deterministic SEA Statistical vibration Network resonator
Wang and Duhamel
0e500
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 39 Summary of mechanism separation models on TPIN.
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
79
4.4.4. Air pumping (Wei et al., 2016) Wei et al. (2016) [233] developed a hybrid numerical-experimental model to predict the air pumping noise at any field point. This model is based on a hand carved tread pattern with two longitudinal grooves and one lateral groove. The groove is meshed into regular triangular or quadrilateral elements, which can be considered as small point acoustic sources. The volume velocity or source strength Q for each element i is given by
Qi ðf Þ ¼ Si $qðf Þ$hðfÞ
(236)
where f indicates frequency domain, S is the surface area of each element, and q is the sound source of unit area that is related to the vehicle speed; h(f) is a coefficient that is a function of f, which is the distance between the element and the leading edge/trailing edge of the contact patch. The velocity of the groove volume in the frequency domain is obtained by the summation of all the elements, as given below.
Qn ¼
X
Qi
(237)
i
where n is the index of sound sources in the groove. In this case, there are four sources: two (near the longitudinal grooves) at the leading edge and two at the trailing edge. The sound pressure spectrum P in the frequency domain is obtained as
Pm ¼
X
Zmn Qn
(238)
n
where m is the index of microphone position in the near field, Zmn is the transfer function between sound pressure Pm and source strength Qn. The reverse manipulation can yield Qn if Pm is experimentally measured and Zmn is numerically determined; this is reason for calling this a hybrid model. 4.5. Summary The summary of the tire-pavement interface models is presented in Table 38. Since the late 1990's, more efforts have been taken to determine the contact force/pressure distribution for TPIN modeling. The pavement texture was also converted to intermediate parameters such as enveloped profile that have high correlation with TPIN. For either near-field or far-field noise, all the models were able to predict the noise spectrum of a broad frequency range, and not just the overall noise level. Vehicle speed was typically included as an input parameter. However, the contact force/pressure spectrum was calculated based on pavement texture while assuming a slick tire tread, which might not be accurate for normal patterned tires. The summary of the mechanism separation models is presented in Table 39. With this type of modeling, it is possible to analyze each noise generation mechanism individually and identify each noise source; this has become popular since the late 2000's. Both structure-borne and airborne noises have been investigated. The output of the models was the noise spectrum in 1/3 octave band or in the narrowband. It is seen that pavement texture was more frequently included than the tire parameters, such as tread pattern. The summary of noise propagation models is presented in Table 40. These models have been used to predict the far-field noise level at a specific distance from the traffic. The input parameters typically included vehicle speed, vehicle type, and some pavement parameters. The environmental parameters influencing sound propagation were more important than in the other type of models and were normally included. 5. Other technical categories The readers can follow the paper structure based on the modeling approach categories (i.e., deterministic, statistical, hybrid, and corresponding sub-categories) to identify the models of interest. More importantly, the tables in each summary section for the three categories follow the same format, which can be considered as an index that the readers can refer to and generate other technical categories. For example, if the readers are interested in TPIN models that reveal the tread pattern effect, “tread pattern” or related key words can be searched for in the column of “Tire Parameter” in the summary tables. Then, the TPIN models on tread pattern effect can be extracted, as presented in Table 41. It can be seen that there are seventeen models related to tread pattern effect. Ten models fall under the category of deterministic models, two under statistical models, and five under hybrid models, which indicates that it is not easy to predict the tread pattern effect on tire noise using statistical methods, mainly because it is not easy to quantify the tread pattern, and a large amount of data for different tread patterns is normally not available for statistical analysis. However, the two statistical methods are both neural computation models, suggesting that neural network might be a feasible approach to correlating the tread pattern with the tire noise. The most popular modeling approach for airborne noise, e.g., air pumping, is CFD (five models) since 21st century. To investigate the tread pattern effect, most of the models reach a very high frequency up to 2000 Hz, because the tread pattern noise normally occurs at higher frequencies. The air pumping and tread impact are the most extensively investigated mechanisms related to the tread pattern. Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
80
No.
Model name
Author
Year
Ref.
Method
Output
Frequency range [Hz]
Mechanism
Driver parameter
Tire parameter
Pavement parameter
Environment parameter
1
CRTN (UK) NMPB (France) Highway
DOT
1988
[210]
Line source
Far field, L10
N/A
Traffic
N/A
1996
[221]
Traffic
2010
[230]
Far field, octave band, power, Lden Pass-by, octave band SPL
125e4000
Dai and Lee
Segmentation of line source Ray acoustics
Type, MPD, slope Slope
N/A
N/A
N/A
Traffic
Wei et al.
2016
[233]
Transfer function
Drum, SPL, narrowband
0e3500
Air pumping, radiation
Speed, vehicle type Speed, vehicle type, acceleration Speed, vehicle type, traffic volume 461 kg, 180 kPa, 80 km/h
2 3
4
Air pumping
N/A N/A
205/55R16, hand cut pattern
Type, porosity, permeability ISO 10844
Atmospheric absorption Temperature, pressure
N/A
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 40 Summary of noise propagation models on TPIN.
No. Model name Author
Year Ref.
Method
Output
1
Monopole
Hayden
1971 [29]
Monopole
2
Unified
1981 [30]
Thin shell
3
Plotkin and Stusnick Heckl
Various mechanisms Tire vibration Bremner et al.
1986 [44]
Various
Sound pressure, N/A pure tone 0e8000 Near field, normalized radiation LDV, vibration, SPL 0e2000
Air pumping CFD 10 Pipe resonance
Far field (7.5 m), 1/3 SEA, wavenumber octave band SPL decomposition De Roo and 2000 [77] FEA, BEM Near field, 1/3 Gerretsen octave band SPL; pass-by, SPL Gagen 1999 [95] CFD Near field, sound pressure/energy Far field, 1/3 octave Kim et al. 2006 [99] CFD, band SPL Kirchhoff integral Chen et al. 2014 [11] CFD Near field, 1/3 octave band SPL 2016 [100] CFD Near/far field, Gautam drum, narrowband and SPL Chandy Fabrizi 2016 [104] CFD, LES, Pass-by/in-groove, aeroacoustics narrowband SPL
11 F-GA
Li et al.
2009 [162] F-GA
Octave band SPL
12 ANN
Che et al.
13 Enhanced SPERoN
Kuijpers and Blokland
4
5
TRIAS
6
Shock wave Hybrid CFD
7
8
CFD
9
14 3D Klein envelopment and Cesbron 15 Tire/road Cao et al. coupling 16 Network Wang Resonator and Duhamel 17 Air pumping Wei et al.
1997 [46]
Frequency range [Hz]
Mechanism
Driver parameter Tire parameter
Pavement Environment Category parameter parameter
Air pumping
Speed
Groove geometry
N/A
N/A
Impact, air pumping
Speed, tire load
Various
Texture
N/A
Various
Rotating, loaded Tread pattern
Roughness N/A
100e4000
Impact
Speed, 40 psi
Tread pattern, materials
Texture
N/A
N/A
Impact, air pumping
N/A
Various, TYDAS
Various, RODAS
N/A
Deterministic FEM & BEM
N/A, time domain 250e8000
Air pumping
Speed
N/A
Deterministic CFD
Air pumping, 100 km/h pipe resonance
Groove angle, N/A geometry Ventilation groove, N/A 165/70SR13
N/A
Deterministic CFD
Airborne
N/A
Tread pattern
N/A
N/A
Deterministic CFD
40 km/h
215/60R16, transverse slots
N/A
N/A
Deterministic CFD
Drum with Safety Walk N/A
N/A
Deterministic CFD
N/A
Statistical
0e2000
1000e10000 Airborne, air pumping 1500e2000
Pipe resonance 80 km/h
205/55R16, longitudinal grooves
125e4000
TPIN
N/A
Tread pattern
2012 [164] GA-BP
DR, octave band SPL 125e4000
TPIN
Size, tread pattern
2003 [183] Multivariate regression analysis
CB, LpA,max, 1/3 octave band SPL
125e5000
TPIN
Speed, tire load, inflation 50-120 km/h
Tread pattern, tire width, rubber hardness
2016 [200] Envelopment, CPX, 1/3 octave band; CB, Lpmax linear regression 2008 [204] N/A SPL, narrowband
250e1250
Belt vibration
65-110 km/h
Tread pattern
0e5000
Various
N/A
Tread pattern
2017 [208] BEM, FEM
Lab, SPL
200e2000
Pipe resonance Stationary
2016 [233] Transfer function
Drum, SPL, narrowband
0e3500
Air pumping, radiation
Tread pattern
461 kg, 205/55R16, hand 180 kPa, 80 km/h cut pattern
Standard N/A surface 20 C Shape factor, contact pressure spectrum, mechanical impedance Texture N/A
Sub-category
Deterministic Conventional physics Deterministic Conventional physics Deterministic Conventional physics Deterministic Conventional physics
Statistical Hybrid
Fuzzy curve fitting Fuzzy curve fitting Tire-pavement interface
Hybrid
Tire-pavement interface
Hybrid
Mechanism separation Mechanism separation
Texture, N/A absorption Smooth N/A
Hybrid
ISO 10844 N/A
Hybrid
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Noise propagation 81
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
Table 41 Summary of TPIN models on tread pattern effect.
No. Model name
Author
Year Ref.
Method
Output
Frequency Mechanism range [Hz]
Driver parameter
Tire parameter
Pavement parameter
Environment Category parameter
1
Unified
1981 [30]
Thin shell
Speed, tire load
Various
Texture
N/A
Deterministic Conventional physics
2D ring
1986 [39]
Tire load
Various
Texture
N/A
3
Various mechanisms Tire vibration
Heckl
1986 [44]
Modal expansion Various
Near field, normalized radiation Near field, SPL
0e8000
2
Plotkin and Stusnick Kung et al.
Tread pattern Tread pattern, materials
N/A
1997 [46]
Rotating, loaded Speed, 40 psi
Roughness
Bremner et al.
LDV, vibration, 0e2000 SPL 100e4000 Impact Far field SEA, (7.5 m), 1/3 wavenumber decomposition octave band SPL Image source Impedance, 100e5000 Pavement absorption absorption
Texture
N/A
Deterministic Conventional physics Deterministic Conventional physics Deterministic Conventional physics
N/A
N/A
N/A
Deterministic Conventional physics
Various, slick 205/60R15 Various, TYDAS
Airflow resistivity, porosity, tortuosity Texture
N/A
Deterministic FEM & BEM
Various, RODAS
N/A
Deterministic FEM & BEM
4
0e400
Impact, air pumping Structureborne Various
Sub-category
5
Porous pavement
rengier Be et al.
1997 [51]
6
3D two-plate
Larsson and Kropp
1992 [25]
FE
Pass-by, SPL
0e2000
Structureborne
Tire load
7
TRIAS
De Roo and Gerretsen
2000 [77]
FEA, BEM
N/A
Impact, air pumping
N/A
8
IFEM
Biermann et al.
2007 [23]
IFEM, AstleyLeis elements
0e1400
Impact
Speed
Slick
Texture
N/A
Deterministic FEM & BEM
9
FEM & BEM
Brinkmeier et al.
2008 [17]
FEA, BEM, ALE
100e850
Structureborne
40-120 km/h
N/A
Texture spectrum
N/A
Deterministic FEM & BEM
10 Road input
Kido et al.
2011 [85]
FE
Near field, 1/3 octave band SPL; pass-by, SPL Far field (1 m), 1/3 octave band SPL Pass-by, 1/3 octave band SPL Axle vibration
Texture
N/A
Deterministic FEM & BEM
Rasmussen
2009 [132]
SRTT
Texture, skewness
N/A
Statistical
Traditional regression
12 Asperity
Fujikawa et al.
2009 [108]
Linear regression, bridging filter Linear regression
Speed, tire load Buick Century, 97 km/h
Patch area
11 Concrete texture
Texture impact OBSI, 1/3 octave 500e5000 Texture 630e1000 impact BTT, OASPL, octave
500e4000 TPIN
80 km/h, 3.2 kN, 185/65R15, 170 kPa rib pattern
N/A
Statistical
Traditional regression
500e5000 TPIN
N/A
SRTT
N/A
Statistical
Traditional regression
SRTT
Wind < Statistical 4 km/h Pavement Statistical temperature N/A Statistical
Traditional regression PCA PCA
N/A
PCA
0e150
13 Pavement core
2011 [138]
Linear regression
OBSI, 1/3 octave, OASIL
14
2014 [141]
Linear regression OLS, PCR
OBSI, OASIL, 500e4000 TPIN pass-by, OASPL OBSI, OASIL N/A TPIN
88 km/h 97 km/h 97 km/h
SRTT
Asperity spacing, asperity height unevenness, MPD Airflow resistance, MPD FN, MPD, AG4 Various
2014 [145,146] PCA Energy
BTT, energy
Speed
Same tire
MTD
2015 [151]
CPX, 1/3 octave 315e3150 TPIN band SPL, OASPL
80 km/h
N/A
Texture, MPD, absorption
15 16 17
Reyes and Harvey FDOT Wayson et al. Ongel Asphalt pavement et al. MTD Zhang et al. Thin surfacing Li et al.
2008 [142]
PCR
0e2000
TPIN
Statistical
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
82
Table 42 Summary of TPIN models on pavement texture effect.
SRTT
Type, ESAL, Age, IRI, SR
30 C
OBSI, 1/3 octave 500e5000 TPIN band, OASIL
Traffic volume
N/A
Age, voids, MPD, thickness
Precipitation Statistical
Transfer function
OBSP, 1/3 octave band
10e4000
TPIN
45 km/h
Passenger and truck
N/A
Hybrid
Crosscorrelation, regression Multivariate regression analysis
OBSP, pass-by, OASPL
22e5500
TPIN
Speed
Slick & normal
Texture, contact pressure spectrum Texture index
N/A
Hybrid
CB, LpA,max, 1/3 125e5000 TPIN octave band SPL
50-120 km/h
Slick, tire width
N/A
Hybrid
2003 [183]
Multivariate regression analysis
CB, LpA,max, 1/3 125e5000 TPIN octave band SPL
50-120 km/h
20 C
Hybrid
Tirepavement interface
Klein and Hamet
2007 [187]
Envelopment, regression
CPB, LpA,max, 1/3 100e5000 TPIN octave band SPL
70, 90, 110 km/h
20 C
Hybrid
Tirepavement interface
25 Road excitation
Rustighi et al.
2008 [192]
FEM, BEM, 2D Gaussian
Vibration, SPL
20-100 km/h
N/A
Hybrid
Tirepavement interface
26 Viscoelastic contact
Dubois et al.
2011 [195]
Macroscale, multi-asperity
Near field, 1/3 octave band SPL
N/A
Hybrid
Tirepavement interface
27 Contact distribution
Dubois et al.
2013 [198]
N/A
Hybrid
28 3D envelopment
Klein and Cesbron
2016 [200]
N/A
Hybrid
29 Tire/road Cao et al. coupling 30 Mechanism Dare decomposition
2008 [204]
Weighted linear regression Envelopment, linear regression N/A
N/A
Hybrid
N/A
Hybrid
Tirepavement interface Tirepavement interface Mechanism separation Mechanism separation
N/A
Hybrid
N/A
Hybrid
Khazanovich 2008 [5] and Izevbekhai
LevenbergMarquardt
OBSI, OASIL
19 Bayesian
Yu and Lu
2013 [173]
Bayesian, MCMC, MI
20 Truck contact
Fong
1998 [177]
21 DGAC texture
Domenichini et al.
1999 [18]
22 SPERoN
Beckenbauer and Kuijpers
2001 [181]
23 Enhanced SPERoN
Kuijpers and Blokland
24 HyRoNE
31 Air pumping separation 32 CRTN (UK)
Winroth et al. DOT
2012 [13]
2013 [205] 1988 [210]
N/A
TPIN
10e400
Texture impact, radiation, interior 315e1000 Texture impact
Shape factor, contact pressure spectrum Shape factor, Tread pattern, tire contact pressure width, rubber spectrum, mechanical hardness impedance Enveloped Michelin texture, Energy absorption, XH1/E3A airflow resistance Smooth Texture excitation
90 km/h
Slick 186/57R15
CPX, 1/3 octave 315e1250 Impact band
50-110 km/h
Slick and Patterned
CPX, 1/3 octave 250e1250 Belt band; CB, Lpmax vibration
65-110 km/h
Tread pattern
N/A Speed
Tread pattern SRTT
50-120 km/h
Slick
Speed, vehicle type
N/A
SPL, 0e5000 Various narrowband Nonlinear least OBSI, 1/3 octave 500e5000 Sidewall, squares band tread band, tangential vibration Curve fitting CPB, 1/3 octave 125e4000 Tire vibration, band SPL air pumping Line source Far field, L10 N/A Traffic
Texture, contact force spectrum Texture, contact force spectrum Texture
Texture, absorption Texture levels, skewness, absorption, friction Texture Type, MPD, slope
Statistical
Fuzzy curve fitting Fuzzy curve fitting Tirepavement interface Tirepavement interface Tirepavement interface T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
Mechanism separation Noise propagation 83
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
97 km/h
18 Pavement aging
84
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
If the readers are interested in a specific model in Table 41, the corresponding section that covers the details of the model can be easily found in the main body of the paper. As another example, the TPIN models on pavement texture effect can be extracted based on the keywords in the column of “Pavement parameter” in the summary tables, as presented in Table 42. It can be seen that there are thirty-two models covering the effect of pavement texture on tire noise: ten under deterministic models, nine under statistical models, and thirteen under hybrid models. It may be noted that the most popular modeling approach to investigate the pavement texture effect is tire-pavement interface (nine models), while there is no model using CFD, which may be because it is not easy for CFD to handle the randomness in the pavement texture. The output of these models is the overall noise level or 1/3 octave band SPL, and only one model (Tire/road coupling, Cao et al., 2008 [204]) is able to predict the noise in narrowband. The structureborne noise, e.g., noise due to pavement texture impact, is the main mechanism investigated. However, because the pavement texture is also related to many other mechanisms (such as air pumping, pavement absorption, and tire vibration), many models considered the tire noise due to pavement texture as a whole, not just the impact noise excited by the pavement texture. The same procedure to cross-reference the models from the summary tables discussed above can be applied to other topics of interest, such as speed effect, temperature effect, noise generation mechanism investigated, modeling technique, and output of the model. 6. Conclusions In this work, the models on the tire-pavement interaction noise (TPIN) were divided into three categories: deterministic models, statistical models, and hybrid models. The deterministic TPIN models include conventional physics models, finite element and boundary element models, and computational fluid dynamics models. They generally give insight into the noise generation and propagation mechanisms. The influence of each specific parameter on TPIN can be numerically analyzed. The conventional physics models typically use dynamic laws to calculate the tire vibration modes; the finite element and boundary element models are capable of obtaining the structure-borne tire noise under arbitrary excitations; the computational fluid dynamics models are intended for the airborne noise analysis. However, a large number of tire parameters are unavoidably needed for deterministic modeling, which is very difficult in practice. In addition, most of the models were limited to the frequency range of below 500 Hz, and few models went through experimental validation. Therefore, the accuracy of these models is in question. The statistical TPIN models include traditional regression models, principal component analysis models, and fuzzy curve fitting models. They generally correlate the noise levels with the influencing parameters using only experimental data without the knowledge of underlying mechanisms. They typically have better accuracy than the deterministic models. The traditional regression analysis, including linear and nonlinear regression, is widely used for models with easy form of equations and having not many variables; the principal component analysis can be used to reduce the number of variables; fuzzy logic has the advantage that no specific form of equations is needed for modeling. However, large amounts of data are needed for statistical modeling, especially when the number of variables is large; this results in high labor cost. In addition, most of the current models included only variables related to the pavement; few models investigated the influence of tire parameters except for some fuzzy curve fitting models. It was also unlikely or impossible to gain insight into the noise generation mechanisms or separate the different mechanisms. The hybrid TPIN models include tire-pavement interface models, mechanism separation models, and noise propagation models. They combine the advantage of the deterministic model and statistical model. The hybrid model sheds light on the noise generation mechanism as a deterministic model, and also has good accuracy as a statistical model. The tire-pavement interface model gives insight into the interaction and excitation between the tire and pavement; the mechanism separation model investigates the noise source individually; the noise propagation model is capable of predicting the TPIN at arbitrary locations. However, the hybrid model tends to be more complex than the deterministic or statistical model. The careful selection of intermediate parameters is often the key to the success of the subsequent statistical correlation procedure. The deterministic models were dominant before the 2000s, especially for conventional physics models. With the development of computing technology, FEM and CFD models are being used frequently thereafter. The use of these models has improved the understanding of TPIN generation mechanisms. Further, partly owing to the computing technology and data techniques, the statistical models have started playing an important role after the 2000s. The application of statistical algorithms to TPIN modeling been has found to be successful, and has led to fairly good prediction accuracy, which is very encouraging. In the 2010s, the hybrid models have become increasingly acceptable among researchers. The authors expect that the hybrid modeling approach will be the future direction in this subject. In addition, the on-board noise measurement techniques are more readily accepted as replacements for pass-by noise measurement techniques, because they are able to obtain noise data for a longer duration and with little environmental influence, which is favorable for the data analysis. As the technology of transducer, data acquisition, and computation advances, narrowband spectrum will be used more frequently for digital signal processing instead of octave, 1/3 octave, or 1/12 octave band; as a result, more precise spectral contents can be analyzed and mechanisms can be better identified. Acknowledgments Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
85
This study (Project Code: MODL-2015-B3-8) has been partially supported by the Center for Tire Research (CenTiRe), an NSF-I/UCRC (Industry/University Cooperative Research Centers) program led by Virginia Tech. The authors hereby wish to thank the project mentors and the members of the industrial advisory board (IAB) of CenTiRe for their kind support and guidance.
References [1] D.B. Thrasher, R.F. Miller, R.G. Bauman, Effect of pavement texture on tire/pavement interaction noise, in: SAE Technical Papers, 1976. [2] R.F. Miller, D.B. Thrasher, Passby tire/pavement interaction noise measurement problems, in: SAE Technical Papers, 1976. [3] M.G. Pottinger, K.D. Marshall, J.M. Lawther, D.B. Thrasher, Review of tire/pavement interaction induced noise and vibration, in: ASTM Special Technical Publication, 1986, pp. 183e287. [4] P.R. Donavan, Assessment of tire/pavement interaction noise under vehicle passby test conditions using sound intensity measurement methods, J. Acoust. Soc. Am. 103 (1998). [5] L. Khazanovich, B.I. Izevbekhai, Implication of time-dependent texture degradation on pavement on board sound intensity patterns in MnROAD test cells, in: Proceedings of the International Noise Conference, Dearborn Michigan July 2008, 2008. [6] A. Syamkumar, K. Aditya, V. Chowdary, Development of mode-wise noise prediction models for the noise generated due to tyre-pavement surface interaction, Adv. Mater. Res. 723 (2013) 50e57. [7] U. Sandberg, J.A. Ejsmont, Tyre/road Noise Reference Book, INFORMEX, Kisa, Sweden; Harg, Sweden, 2002. [8] U. Sandberg, J.A. Ejsmont, The Tire/Road noise computerized bibliography, in: International Symposium, Euro Noise ’95: Software for Noise Control; 1995; Lyon; France, 1995, pp. 289e294. [9] P. Sabiniarz, W. Kropp, A waveguide finite element aided analysis of the wave field on a stationary tyre, not in contact with the ground, J. Sound. Vib. 329 (15) (2010) 3041e3064. [10] M. Harrison, Vehicle Refinement: Controlling Noise and Vibration in Road Vehicles, Elsevier, MA, USA, 2004. [11] C. Chen, Y. Kuan, C. Chen, M. Sung, Using CFD technique to investigate the effect of tire roiling-noise with different pattern design, in: Applied Mechanics and Materials, vol. 575, 2014, pp. 469e472. [12] Z. Mohamed, X. Wang, R. Jazar, A survey of wheel tyre cavity resonance noise, Int. J. Veh. Noise Vib. 9 (3e4) (2013) 276e293. [13] T.P. Dare, Generation Mechanisms of Tire-pavement Noise, Purdue University, 2012. [14] A. Kuijpers, G. van Blokland, Tyre/road noise models in the last two decades: a critical evaluation, in: Proceedings of Inter-Noise 2001, The Hague, Holland, 2001, p. 2494. [15] M.J. Crocker, Handbook of Noise and Vibration Control, John Wiley and Sons Inc., USA, 2007. [16] T. Li, Tire-pavement Interaction Noise (TPIN) Modeling Using Artificial Neural Network (ANN), Appendix H: a Literature Review of Tire Pavement Interaction Noise, PhD Dissertation, Virginia Tech, 2017. [17] M. Brinkmeier, U. Nackenhorst, S. Petersen, O. von Estorff, A finite element approach for the simulation of tire rolling noise, J. Sound. Vib. 309 (1e2) (Jan. 2008) 20e39. , Relationship between road surface characteristics and noise [18] L. Domenichini, A. Fracassa, F. La Torre, G. Loprencipe, A. Ranzo, A. Scalamandre emission, in: First International Colloquium on Vehicle Tyre Road Interaction, Rome, Italy, Paper 99.03, 1999, pp. 1e22. [19] Z. Lou, An Experimental and Numerical Study of Tire/pavement Noise on Asphalt Rubber Concrete Pavement, University of Regina, 2007. [20] G. Wang, H. Zhou, Z. Mao, L. Gao, Boundary element analysis of rolling tire noise, in: Proceedings 2011 International Conference on Transportation, Mechanical, and Electrical Engineering, TMEE 2011, 2011, pp. 1970e1973. [21] Y. Nakajima, Y. Inoue, H. Ogawa, Application of the boundary element method and modal analysis to tire acoustics problems, Tire Sci. Technol. 21 (2) (1993) 66e90. [22] J.S. Yang, T.F. Fwa, G.P. Ong, C.H. Chew, Finite-element analysis of effect of wide-base tire on tire-pavement noise, Adv. Mater. Res. 723 (2013) 105e112. [23] J. Biermann, O. von Estorff, S. Petersen, H. Schmidt, Computational model to investigate the sound radiation from rolling tires, Tire Sci. Technol. 35 (3) (2007) 209e225. [24] K. Larsson, W. Kropp, A high-frequency three-dimensional tyre model based on two coupled elastic layers, J. Sound. Vib. 253 (4) (2002) 889e908. [25] W. Kropp, Ein Modell zur Beschreibung des Rollgera€usches eines unprofiliertenGu€rtelreifens auf rauher Strabenoberfla€che (A model for describing the rolling noise of an unprofiled belt tire on a rough road surface), in: Thesis, Fortschritt-berichte Reihe 11 (Progress Reports Series 11), Nr 166, VDI Verlag, Du€sseldorf, 1992. [26] K. Larsson, W. Kropp, A high frequency range tyre model based on two coupled elastic plates, in: Proceedings of INTER-NOISE 99, Fort Lauderdale, FL, 1999, pp. 131e136. [27] P. Andersson, K. Larsson, W. Kropp, A method for experimental collection of global material data for tyres, in: Conference Nordic Vibration Research 2001, Stockholm, Sweden, 2001. [28] P. Andersson, K. Larsson, Validation of a high frequency three-dimensional tyre model, Acta Acust. united Acust. 91 (1) (2005) 121e131. [29] R.E. Hayden, Roadside noise from the interaction of a rolling tire with the road surface, in: Proceedings of the Purdue Noise Control Conference, Purdue University, Indiana, USA, 1971. [30] K.J. Plotkin, E. Stusnick, A unified set of models for tire/road noise generation, in: Report: WR-81-26, 63p, Environmental Protection Agency, Arlington, VA, Office of Noise Abatement and Control, 1981, p. 63. [31] F. Bohm, Zur Statik und Dynamik Des Gurtelreifens (On the statics and dynamics of the belt tightening), ATZ 69 (1967) 225e261. [32] K.J. Plotkin, M.L. Montroll, W.R. Fuller, A study to indentify advanced methods of reducing tire noise, in: Wyle Reserach Report WR 80-16, for the U.S. Environmental Protection Agency, 1980. [33] N.A. Nilsson, Generation mechanism of external tire noise, in: IFM Akustikbyran Report No. TR3.709.14, 1976. [34] A.C. Eberhardt, Investigations of the truck tire vibration sound mechanisms, in: Paper Presented at the International Tire Noise Conference, Stockholm, 1979. [35] K.J. Plotkin, M.L. Montroll, W.R. Fuller, A study to identify advanced methods of reducing tire noise, in: Wyle Reserach Report WR 80-16, April 1980, Wyle Laboratories, Arlington, Virginia, 1980. [36] S. Samuels, Recent australian tyre/road noise research, in: Proc. Of the International Tire Noise Conference, STU-information No. 168-1980, NUTEK, Stockholm, 1979, pp. 1e25. [37] P.M. Morse, Theoretical acoustics, Am. J. Phys. 38 (5) (1970) 666. [38] R.F. Keltie, Analytical model of the truck tire vibration sound mechanism, J. Acoust. Soc. Am. 71 (2) (1982) 359. [39] L.E. Kung, W. Soedel, T.Y. Yang, Free vibration of a pneumatic tire-wheel unit using a ring on an elastic foundation and a finite element model, J. Sound. Vib. 107 (2) (1986) 181e194. [40] S.C. Huang, W. Soedel, Response of rotating rings to harmonic and periodic loading and comparison with the inverted problem, J. Sound. Vib. 118 (2) (1987) 255e270. [41] S.C. Huang, C.K. Su, In-plane dynamics of tires on the road based on an experimentally verified ring model, Veh. Syst. Dyn. 21 (1992) 247e267. [42] R.L. Wheeler, H.R. Dorfi, B.B. Keum, Vibration modes of radial tires: measurement, prediction, and categorization under different boundary and operating conditions, in: SAE Technical Paper 2005-01-2523, 2005.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
86
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
[43] F. Boehm, Mechanik des Gürtelreifens (Mechanics of the radial tire), Ing.-Arch 35 (1966) 82. [44] M. Heckl, Tire noise generation, Wear 113 (1986) 157e170. €uschentstehung und -abstrahlung [45] D. Ronneberger, Experimentelle und Theoretische Untersuchungen Spezieller Mechanismen der Rollgera (Experimental and theoretical studies specifically mechanisms of rolling noise generation and radiation), in: Mitteilung Nr. 57, Institut für Strassenund Eisenbahn- und Felsbau and der E.T.H. Zürich, 1984, p. 79. [46] P. Bremner, J. Huff, J.S. Bolton, A model study of how tire construction and materials affect vibration-radiated noise, in: SAE Technical Papers, 1997. [47] M. Mikulas, McElman, On free vibrations of eccentrically stiffened cylindrical shells flat plates, in: NASA TN D-3010, 1965. [48] D.C. Rennison, Interior noise control prediction study for high-speed propeller driven aircraft, in: NASA Contractor Report, CR159200, 1979. [49] S. Saigal, T.Y. Yang, Free vibrations of a tire as a toroidal membrane, J. Sound. Vib. 107 (1) (1986) 71e82. [50] P.J. Remington, N.R. Dixon, L.E. Wittig, L.G. Kurzweily, C.W. Menge, J.D. Stahr, Control of wheel/rail noise and vibration, in: US DOT Report DOT-TSCUMTA-82-57, 1983. rengier, M.R. Stinson, G.A. Daigle, J.F. Hamet, Porous road pavements: acoustical characterization and propagation effects, J. Acoust. Soc. Am. [51] M.C. Be 101 (1) (1997) 155. [52] J.S. Bolton, H.J. Song, Y.K. Kim, Y.J. Kang, Wave number decomposition approach to the analysis of tire vibration, in: Proceedings - National Conference on Noise Control Engineering, 1998, pp. 97e102. [53] W. Kropp, A Mathematical model of tyre noise generation, Heavy Veh. Syst. 6 (1e4) (1999) 310e329. €rpers in Luft mit beliebiger Schnelleverteilung aus Kugelschallfeldern (The [54] L. Cremer, Die Synthese des Schallfeldes eines beliebigen festen Ko synthesis of the sound field of any solid body in air with any velocity distribution from spherical sound fields), Acustica 55 (1984) 44. [55] M. Heckl, Bemerkungen zur Berechnung der Schallabstrahlung nach der Methode der Kugelfeldsynthese (Remarks on the calculation of the sound radiation by the method of spherical field synthesis), Acustica 68 (1989) 251. [56] M.F. Übler, Calculation of tyre motions with the folded plate model and comparison with measurements, in: Poster Presented at the Transport Noise Conference, St. Petersburg, Russia, 1994. [57] Y.-J. Kim, J.S. Bolton, Effect of rotation on the vibration characteristics of tires, in: InterNoise 2002, the 2002 International Congress and Exposition on Noise Control Engineering Dearborn, MI, USA, 2002. [58] Y.J. Kim, J.S. Bolton, Effects of rotation on the dynamics of a circular cylindrical shell with application to tire vibration, J. Sound Vib. 275 (3e5) (2004) 605e621. [59] M. Matsubara, N. Tsujiuchi, T. Koizumi, A. Ito, K. Bito, Natural frequency analysis of tire vibration using a thin cylindrical shell model, in: SAE Tech. Pap. 2015-01-2198, 2015. [60] L.R. Molisani, R.A. Burdisso, D. Tsihlas, A coupled tire structure/acoustic cavity model, Int. J. Solids Struct. 40 (19) (2003) 5125e5138. [61] T.L. Richards, Finite element analysis of structural-acoustic coupling in tyres, J. Sound. Vib. 149 (2) (1991) 235e243. [62] B.S. Kim, G.J. Kim, T.K. Lee, The identification of sound generating mechanisms of tyres, Appl. Acoust. 68 (1) (2007) 114e133. [63] P. Gu, Z.C. Feng, Y. Chen, Z. Li, P. Gu, Y. Chen, Z. Li, Modeling and experimental investigation of tire cavity noise generation mechanisms for a rolling tire, SAE Int. J. Passeng. Cars - Mech. Syst. 2 (1) (2009) 1414e1423. [64] Z.C. Feng, P. Gu, Modeling and experimental verification of vibration and noise caused by the cavity modes of a rolling tire under static loading, in: SAE Technical Paper 2011-01-1581, 2011. [65] D.J. O'Boy, A.P. Dowling, Tyre/road interaction noise-A 3D viscoelastic multilayer model of a tyre belt, J. Sound Vib. 322 (4e5) (2009) 829e850. [66] D.J. O'Boy, A.P. Dowling, Tyre/road interaction noise-Numerical noise prediction of a patterned tyre on a rough road surface, J. Sound. Vib. 323 (1e2) (2009) 270e291. [67] K.F. Graff, Wave Motion in Elastic Solids, 1991. [68] I. Lopez Arteaga, Green's functions for a loaded rolling tyre, Int. J. Solids Struct. 48 (25e26) (2011) 3462e3470. [69] I. Lopez, R.E.A. Blom, N.B. Roozen, H. Nijmeijer, Modelling vibrations on deformed rolling tyres-a modal approach, J. Sound. Vib. 307 (3e5) (2007) 481e494. [70] J.S. Bolton, R. Cao, Effect of rotation on the natural frequencies of coupled tire structural-acoustical mode, in: Publications of the Ray W. Herrick Laboratories. Paper 94, 2013. [71] R. Cao, J.S. Bolton, Improved model for coupled structural-acoustic modes of tires, SAE Int. J. Passeng. Cars - Mech. Syst. 8 (3) (2015) 845e854. [72] J.S. Bolton, R. Cao, Point excitation of a coupled structural-acoustical tire model with experimental verification, in: Publications of the Ray W. Herrick Laboratories. Paper 111, 2015. [73] C.G. Diaz, P. Kindt, J. Middelberg, S. Vercammen, C. Thiry, R. Close, J. Leyssens, Dynamic behaviour of a rolling tyre: experimental and numerical analyses, J. Sound. Vib. 364 (2016) 147e164. [74] C. Wright, G.H. Koopmann, A technique to predict the acoustic radiation characteristics of an automobile tire, Tire Sci. Technol. 14 (2) (1986) 102e115. [75] T.L. Richards, Finite Element Analysis of Cord-rubber Structures in Cylindrical Coordinates, 1987. Firestone Tire and Rubber Co., Internal Report. [76] T. Berge, S.A. Storeheier, Tyre/road noise and road texture - a comparison study of modelling and measurements results, in: International Congress on Noise Control Engineering 2005, INTERNOISE 2005, vol. 5, 2005, pp. 4016e4022. [77] F. De Roo, E. Gerretsen, TRIAS - tyre road interaction acoustic simulation model, in: InterNoise 2000, The 29th International Congress and Exhibition on Noise Control Engineering 27-30 August 2000, Nice, FRANCE, vol. 4, 2000, pp. 2488e2496. [78] F. de Roo, E. Gerretsen, W.G.J. Hoffmans, N.J. Doelman, Dutch tyre-road noise emission model - adjustments and validation, in: Proceedings Internoise ’99, Fort Lauderdale, Florida, USA, 1999, pp. 243e248. [79] C.-M. Nilsson, Waveguide Finite Elements Applied on a Car Tyre, Doctoral thesis, Department of Aeronautical and Vehicle Engineering, Royal Institute of Technology, 2004. [80] M. Fraggstedt, Power Dissipation in Car Tyres, Licentiate thesis, Department of Aeronautical and Vehicle Engineering, Royal Institute of Technology, 2006. [81] C. Hoever, A. Tsotras, E.-U. Saemann, W. Kropp, A comparison between finite element and waveguide finite element methods for the simulation of tire/road interaction, in: 42nd International Congress and Exposition on Noise Control Engineering 2013, INTER-NOISE 2013: Noise Control for Quality of Life, vol. 1, 2013, pp. 127e136. [82] Y. Waki, K. Kozuki, H. Heguri, Estimation of noise radiating parts of a tire using the wave finite element method, in: 40th International Congress and Exposition on Noise Control Engineering 2011, INTER-NOISE 2011, vol. 1, 2011, pp. 577e584. [83] Y. Waki, B.R. Mace, M.J. Brennan, Free and forced vibrations of a tyre using a wave/finite element approach, J. Sound. Vib. 323 (2009) 737e756. [84] C.G. Diaz, S. Vercammen, J. Middelberg, C.T.P. Kindt, J. Leyssens, Numerical prediction of the dynamic behaviour of rolling tyres, in: Proceedings of ISMA 2012, 2012, pp. 1541e1551. [85] I. Kido, S. Ueyama, M. Hashioka, S. Yamamoto, M. Tsuchiyama, H. Yamaoka, Tire and road input modeling for low-frequency road noise prediction, SAE Tech. Pap. 4 (2) (2011) 1277e1282. [86] C. Wagner, T. Huttl, P. Sagaut, Large-eddy Simulation for Acoustics, 2007. [87] U. Sandberg, J.A. Ejsmont, E. Gustavsson, Tire/road Noise on Rubberized Asphalt and Cement Concrete Surfaces in Sweden, 1990, p. 60. [88] PIARC, Specification for a Standard Test Tire for Friction Coefficient Measurement of a Pavement Surface: Smooth Test Tire, World Road Assoc., 2004. [89] G. Tong, X. Jin, W. Tian, K. Yang, Simulation on factors affecting radial tire noise, J. Converg. Inf. Technol. South. Korea) 8 (8) (2013) 209e216. [90] N. Vlahopoulos, G. Zhang, W. Brophy, M. Ramaswami, Vehicle airborne noise analysis using the energy finite element method, SAE Tech. Pap. 4 (2013). [91] Z. Mohamed, X. Wang, A study of tyre cavity resonance and noise reduction using inner trim, Mech. Syst. Signal Process. 50e51 (2015) 498e509. [92] R.D. Blevins, Formulas for Natural Frequency and Mode Shapes, Krieger Pub Co, Florida, 1995.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
87
[93] Y. Wei, X. Feng, Z. Fuqiang, D. Xiang, Simulation of rolling noise based on the mixed Lagrangian-Eulerian method, Tire Sci. Technol. 44 (1) (2016) 36e50. [94] L. Banz, H. Gimperlein, Z. Nezhi, E.P. Stephan, Time domain BEM for sound radiation of tires, Comput. Mech. 58 (1) (2016) 45e57. [95] M.J. Gagen, Novel acoustic sources from squeezed cavities in car tires, J. Acoust. Soc. Am. 106 (2) (1999) 794e801. [96] M.J. Gagen, Nonlinear acoustic sources in squeezed car tyre cavities, Noise Vib. Worldw. 31 (4) (2000) 9e19. [97] L.E. Kinsler, A.R. Frey, A.B. Coppens, J.V. Sanders, Fundamentals of Acoustics, fourth ed., Wiley-VCH, 1999, p. 560. ISBN 0-471-84789-5. [98] J.A. Ejsmont, U. Sandberg, S. Taryma, Influence of tread pattern on tire/road noise, in: Passenger Car Meeting, October 1, 1984-October 4, 1984, 1984, p. 9. [99] S. Kim, W. Jeong, Y. Park, S. Lee, Prediction method for tire air-pumping noise using a hybrid technique, J. Acoust. Soc. Am. 119 (6) (2006) 3799e3812. [100] P. Gautam, A.J. Chandy, Numerical investigation of the air pumping noise generation mechanism in tire grooves, J. Vib. Acoust. Trans. ASME, Pap. No VIB-15-1493 138 (5) (2016), 51002e51002e8. [101] P. Gautam, A.J. Chandy, A three-dimensional numerical investigation of air pumping noise generation in tires, J. Vib. Acoust. Trans. ASME 138 (6) (2016). [102] K. Takami, T. Furukawa, Study of tire noise characteristics with high-resolution synchronous images, in: EURONOISE, Maastricht, The Netherlands, May 31eJune 3, 2015, pp. 2113e2118. [103] K. Takami, T. Furukawa, High-resolution deformation measurement system for fast rotating tires, in: Proceedings of the ASME Design Engineering Technical Conference vol. 8, 2015, p. 6. [104] C. Fabrizi, Computational aeroacoustic analysis of a rolling tire, Tire Sci. Technol. 44 (4) (Oct. 2016) 262e279. [105] M.J. Lighthill, On sound generated aerodynamically. I. General theory, Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 211 (1952) 565e587. [106] M.J. Lighthill, On sound generated aerodynamically. II. Turbulence as a source of sound, Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 222 (1954) 1e32. [107] R. Rasmussen, S. Karamihas, E. Mun, G. Chang, Relating pavement texture to tire-pavement noise, in: Institute of Noise Control Engineering of the USA - 35th International Congress and Exposition on Noise Control Engineering, INTER-NOISE 2006, vol. 1, 2006, pp. 422e431. [108] T. Fujikawa, Y. Oshino, T. Mikami, H. Tachibana, Examination of road roughness parameters for abating tire vibration and radiated noise, Noise Control Eng. J. 57 (2) (2009) 77e83. [109] J. Quartieri, N.E. Mastorakis, G. Iannone, C. Guarnaccia, S. D'Ambrosio, A. Troisi, T.L.L. Lenza, A review of traffic noise predictive models, in: Proceedings of the 5th WSEAS Int. Conf. On “Applied and Theoretical Mechanics”(MECHANICS’09), Puerto De La Cruz, Canary Islands, Spain, 2009, pp. 72e80. [110] Anon, Handbook of Acoustic, Noise control WADC technical report, 52-204, Wright Air Development Center, 1952. [111] A. Nickson, Can community reaction to increased traffic noise be forecast?, in: Proc. Fifth International Congress on Acoustics, 1965. [112] D. Johnson, E. Saunders, The evaluation of noise from freely flowing road traffic, J. Sound. Vib. 7 (2) (1968) 287e309. [113] I. Griffiths, F. Langdon, Subjective Response to road traffic noise, J. Sound. Vib. 8 (1968) 16e32. [114] W. Galloway, Urban Highway Noise: Measurement, Simulation and Mixed Reactions vol. 78, 1969. NCHRP report. [115] M.A. Burgess, Noise prediction for urban traffic conditions. Related to measurement in sydney metropolitan area, Appl. Acoust. 10 (1977) 1e7. [116] Centre Scientifique et Technique du Batiment (Scientific and Technical Center of the Building), Etude theorique et experimentale de la propagation acoustique (Theoretical and experimental study of acoustic propagation), in: Revue D'Acoustique n.70 (Acoustic Review n.70), 1991. [117] C. Fagotti, A. Poggi, Traffic noise abatement strategies. The analysis of real case not really effective, in: Proc. Of 18th International Congress for Noise Abatement, pp., Bologna, Italy, 1995, pp. 223e233. [118] E. Murphy, E.A. King, Environmental Noise Pollution, 1st Edition: Noise Mapping, Public Health, and Policy, Elsevier, 2014. [119] Der Bundesminister für Verkehr (The Federal Minister of Transport), Richtlinien fur den Larmschutz an Strassen (Guidelines for the noise protection on roads, in: RLS81, 1981. [120] RLS, Richtlinien fu€r den La€rmschutz an Strassen (Guidelines for noise protection on roads), in: BM Fu€r Verkehr (The Federal Minister of Transport), Bonn, 1990. [121] G. B. Canelli, K. Gluck, and S. A. Santoboni, “A mathematical model for evaluation and prediction of mean energy level of traffic noise in Italian towns,” Acustica, vol. 53, no. 31, p. 1983. [122] A. Cocchi, A. Farina, G. Lopes, Modelli matematici per la previsione del rumore stradale: verifica ed affinazione del modello CNR in base a rilievi sperimentali nella citta di Bologna (Mathematical models for road noise prediction: verification and refinement of the CNR model based on, in: Acta of 19 Convegno Nazionale AIA (Acta of 19th National AIA Conference), 1991. [123] Federal Highway Authority, Code of Federal Regulations, Title 23 e Highways Part, Section 772.9 e Traffic Noise Prediction, US Dep. Transp., 2012. [124] Federal Highway Authority, FHWA Traffic Noise Model Users Guide, US Department of Transportation, 1998. Report N. FHWA-PD-96-009 and DOTVNTSC-FHWA-98-1. [125] G. Fleming, A. Rapoza, C. Lee, Development of national reference energy mean emis- sion levels for the FHWA Traffic Noise Model, in: Report No FHWA-pd-96-008 and DOT- VNTSC-96-2 John A, Volpe National Transportation Systems Center, 1995. [126] FHWA, FHWA Traffic Noise Model, 2013, Version 3.0. http://www.fhwa.dot.gov/environment/noise/traffic_noise_model/tnm_v30/. [127] J.L. Rochat, A.L. Hastings, D.R. Read, M. Lau, FHWA Traffic Noise Model ( TNM ) Pavement Effects Implementation Study : Progress Report 1, 2012. Report #: FHWA-HEP-12-034. [128] D.I. Hanson, R.S. James, Colorado DOT Tire/Pavement Noise Study, Colorado Department of Transportation, 2004. Report No. CDOT-DTD-R-2004-5. [129] D.I. Hanson, R.S. James C, C. NeSmith, Tire/pavement Noise Study, 2004, p. 49. NCAT Report 04-02. [130] TXDOT, Fineness modulus of fine aggregate, in: TxDOT Designation: Tex-402-A, 1999. [131] ASTM, Method of Calculation of the Fineness Modulus of Aggregate, 1980. CRD-C 104. ́ ́ J. 57 (2) (2009) 139e147. [132] R. Rasmussen, Measuring and modeling tire-pavement noise on various concrete pavement textures, Noise Control Eng. [133] R.O. Rasmussen, Quieter concrete pavements - an update of pooled fund TPF-5(139), in: Presented at Pavement Evaluation 2010, Roanoke, Virginia, vol. 5, 2010 no. October. [134] ISO, Characterization of Pavement Texture by Use of Surface Profiles e Part 4: Spectral Analysis of Surface Profiles, 2008. ISO/TS 13473-4. [135] ASME, Surface Texture (Surface Roughness, Waviness, and Lay), 2009. ASME B46.1. [136] ISO, Geometrical Product Specifications (GPS) e Surface Texture: Profile Method; Surfaces Having Stratified Functional Properties e Part 2: Height Characterization Using the Linear Material Ratio Curve, 1996. ISO 13565-2. [137] T. Fujikawa, Y. Oshino, T. Mikami, H. Tachibana, Examination on effects of road roughness parameters for abating tire/road noise, in: Proceedings of Inter-noise 2004, Prague, Czech Republic, 2004. [138] C.H. Reyes, J. Harvey, A method for predicting sound intensity noise levels using laboratory pavement cores, in: Proceedings of Noise-con 2011, Portland, OR, 2011, pp. 522e534. [139] ISO, Characterization of Pavement Texture by Use of Surface Profiles e Part 1: Determination of Mean Profile Depth, 1997. ISO 13473-1. ̀ [140] S. Kephalopoulos, M. Paviotti, F. Anfosso-Ledee, Common Noise Assessment Methods in Europe (CNOSSOS-EU), European Commission Joint Research Centre, Luxembourg publications Office of the European Union, 2012. [141] L.R. Wayson, M.J. MacDonaldA, A. Martin, On-board Sound Intensity (OBSI) Study: Phase 2, 2014, p. 145. Final Report, FDOT Project No. #BDT06. [142] A. Ongel, E. Kohler, J. Harvey, Principal components regression of onboard sound intensity levels, J. Transp. Eng. 134 (11) (2008) 459e466. [143] AASHTO, Standard Method of Test for Measurement of Tire/Pavement Noise Using the On-board Sound Intensity (OBSI) Method, 2008. AASHTO TP 76. [144] A. Ongel, J. Harvey, Pavement characteristics affecting the frequency content of tire/pavement noise, Noise Control Eng. J. 58 (6) (Nov. 2010) 563e571.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
88
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
[145] Y. Zhang, J.G. McDaniel, M.L. Wang, Estimation of pavement macrotexture by principal component analysis of acoustic measurements, J. Transp. Eng. 140 (2) (2014) 04013004 (11 pp.). [146] Y. Zhang, J.G. McDaniel, M.L. Wang, Pavement macrotexture estimation using principal component analysis of tire/road noise, in: Proc. SPIE, Int. Soc. Opt. Eng. (USA), vol. 9063, 2014, p. 90630K (9 pp.). [147] Y. Zhang, X. Ma, J.G. McDaniel, M.L. Wang, Statistical analysis of acoustic measurements for assessing pavement surface condition, in: Proc. Of SPIE Smart Structures and Materialsþ Nondestructive Evaluation and Health Monitoring (Pp. 83471F-83471F), International Society for Optics and Photonics, Bellingham, WA, 2012. [148] H.Q. Wang, Z.H. Song, H. Wang, Statistical process monitoring using improved PCA with optimized sensor locations, J. Process. Contr. 12 (6) (2002) 735e744. [149] H. Hotelling, Analysis of a complex of statistical variables into principal components, J. Educ. Psychol. 24 (6) (1933), 417e441e520. [150] I.T. Jolliffe, Principal Component Analysis, second ed., Springer-Verlag, New York, 2002. [151] M. Li, W. van Keulen, H. Ceylan, G. Tang, M. van de Ven, A. Molenaar, Influence of road surface characteristics on tire-road noise for thin-layer surfacings, J. Transp. Eng. 141 (11) (2015). [152] S. Roychowdhury, Fuzzy curve fitting using least square principles, in: Proceedings of the IEEE International Conference on Systems, Man and Cybernetics, vol. 4, 1998, pp. 4023e4027. [153] J. Fry, P. Jennings, N. Taylor, P. Jackson, Vehicle Drive-by Noise Prediction: a Neural Networks Approach, 1999. SAE Technical Paper Series 1999-011740. [154] J. Fry, P. Jennings, Using multi-layer perceptrons to predict vehicle pass-by noise, Neural Comput. Appl. 11 (3) (2003) 161e167. [155] ISO, Measurement of Noise Emitted by Accelerating Road Vehicles e Engineering Method e Part 1: M and N Categories, 2015. ISO 362-1. [156] ISO, Measurement of Noise Emitted by Accelerating Road Vehicles e Engineering Method e Part 2: L Category, 2009. ISO 362-2. [157] L. Tarassenko, Guide to Neural Computing Applications, Butterworth-Heinemann, 1998. [158] P.D. Wasserman, Advanced Methods in Neural Computing, Van Nostrand Reinhold, New York, 1993. [159] R.N. Neal, Bayesian Learning for Neural Networks, Springer-Verlag, 1996. [160] ASTM, Standard Specification for P225/60R16 97S Radial Standard Reference Test Tire, 2014. ASTM F2493. [161] B.I. Izevbekhai, Tire Pavement Interaction Noise of Concrete Pavements, PhD Dissertation, University of Minnesota, 2012. [162] X.-H. Li, J. Liu, D.-Q. Liu, K.-M. Sun, Application of tread patterns noise-reduction based on self-adaptive fuzzy genetic algorithm, Comput. Eng. Appl. 45 (22) (2009). [163] U. Kumar, J. Prakash, G. Unnikrishnan, T. Thomas, Tire pitch sequence optimization using HyperStudy, in: Altair Technology Conference, India, 2013. [164] Y. Che, W. Xiao, L. Chen, Z. Huang, C. Yong, X. Wangxin, C. Lijun, H. Zhichu, GA-BP neural network based tire noise prediction, Adv. Mater. Res. 443e444 (2012) 65e70. [165] C. Yong, X. Wangxin, C. Lijun, Laboratory test and spectrum analysis method of tire noise, in: 2009 International Conference on Measuring Technology and Mechatronics Automation (ICMTMA), vol. 1, IEEE Press, Apr. 2009, pp. 487e490, 2009. [166] P. Kostial, Z. Jancikova, D. Bakosova, J. Valicek, M. Harnicarova, I. Spicka, P. Kotial, D. Bakoova, I. picka, Artificial neural networks application in modal analysis of tires, Meas. Sci. Rev. 13 (5) (2013) 273e278. [167] W. Kongrattanaprasert, H. Nomura, T. Kamakura, K. Ueda, Application of neural network analysis to automatic detection of road surface conditions utilizing tire noise from vehicles, in: 2009 ICROS-SICE International Joint Conference. ICCAS-SICE 2009, 2009, pp. 2354e2358. [168] W. Kongrattanaprasert, H. Nomura, T. Kamamura, K. Ueda, Automatic detection of road surface states from tire noise using neural network analysis, in: 20th International Congress on Acoustics 2010, ICA 2010-Incorporating Proceedings of the 2010 Annual Conference of the Australian Acoustical Society vol. 4, 2010, pp. 2965e2968. [169] M. Kandpal, V.K. Kakar, G. Verma, Classification of ground vehicles using acoustic signal processing and neural network classifier, in: 2013 International Conference on Signal Processing and Communication, ICSC 2013, 2013, pp. 512e518. [170] Y. Nakajima, H. Kadowaki, T. Kamegawa, K. Ueno, Application of a neural network for the optimization of tire design, Tire Sci. Technol. 27 (2) (1999) 62e83. [171] M. El-Gindy, H.L. Lewis, A.S. Lewis, Development of a tire/pavement contact-stress model based on an artificial neural network, Am. Soc. Mech. Eng. Des. Eng. Div. 101 (1999) 25e34. [172] M. El-Gindy, H. Lewis, Development of a Tire/Pavement Contact-stress Model Based on Artificial Neural Networks, 2001. Final Report No. FHWA-RD99-041. [173] B. Yu, Q. Lu, Bayesian model for tyre/asphalt pavement noise, Proc. Inst. Civ. Eng. - Transp. 166 (4) (2013) 241e252. [174] Q. Lu, E. Kohler, J. Harvey, A. Ongel, Investigation of Noise and Durability Performance Trends for Asphaltic Pavement Surface Types: three-Year Reports, University of California, Davis, CA, USA, 2009. Report No. UCPRC-RR- 2009e01. [175] Q. Lu, E. Kohler, J. Harvey, Field investigation of acoustic performance of various asphalt surface mixes, in: ICCTP 2010: Integrated Transportation Systems: Green, Intelligent, Reliable - Proceedings of the 10th International Conference of Chinese Transportation Professionals, vol. 382, 2010, pp. 3937e3947. [176] ASTM, Standard Practice for Calculating Pavement Macrotexture Mean Profile Depth, 2009. ASTM E1845. [177] S. Fong, Tyre noise predictions from computed road surface texture induced contact pressure, in: Proceedings of INTER-NOISE 98 Christchurch, New Zealand, 1998, pp. 137e140. [178] A. D'Andrea, TINO prototype of a low noise asphalt, in: First International Colloquium on Vehicle Tyre Road Interaction- the Noise Emission, Rome, Italy, 1999. [179] ISO, Characterization of Pavement Texture by Use of Surface Profiles e Part 2: Terminology and Basic Requirements Related to Pavement Texture Profile Analysis, 2002. ISO 13473-2. [180] S. Huschek, Characterization of pavement surface texture and its influence on tire/road noise, in: Proceedings of the Third International Symposium on Pavement Surface Characteristics: Christchurch, New Zealand 3-4 September 1996, 1996. [181] T. Beckenbauer, A. Kuijpers, Prediction of pass-by levels depending on road surface parameters by means of a hybrid model, in: Proceedings of Internoise 2001, the Hague, Holland, 2001, pp. 2528e2533. [182] T.G. Clapp, A.C. Eberhardt, C.T. Kelley, Development and validation of a mehod for approximating road surface texture-induced contact pressure in tire-pavement interaction, Tire Sci. Technol. 16 (1) (1988) 2e17. [183] A. Kuijpers, G. van Blokland, Tyre/road noise modeling: the road from a tyre's point-of-view, in: Proceedings of Inter-noise 2003, the 32nd International Congress and Exposition on Noise Control Engineering Jeju International Convention Center, Seogwipo, Korea, August 25-28, 2003, 2003, p. 249. cot, Thomas Beckenbauer, The modelling of tyre/road noise d a quasi three-dimensional [184] W. Kropp, K. Larsson, F. Wullens, P. Andersson, F.-X. Be model, in: InterNoise2001, the Hague, The Netherlands, 2001, p. 657. [185] G. van Blokland, P. The, Test sections for development of a hybrid tyre/road interaction noise model, in: Proceedings of Inter-noise 2007, Istanbul, Turkey, 2007. [186] T. Beckenbauer, P. Klein, J.-F. Hamet, W. Kropp, “Tyre/road noise prediction: a comparison between the SPERoN and HyRoNE models - Part 1, in: Proceedings of Acoustics08, Paris, France, 2008, pp. 2933e2938. [187] P. Klein, J.-F. Hamet, Tyre/road noise prediction with the HyRoNE model, in: Proceedings of INTER-NOISE 07, Istanbul, Turkey, 2007, p. 191. [188] J.F. Hamet, P. Klein, Evaluation of tire road noise from road texture profiles using INRETS rolling tire model: evaluation from a profile measured by the LCPC, in: Rapport de recherche, 2004, p. 18.
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and Vibration (2018), https://doi.org/10.1016/j.jsv.2018.01.026
T. Li et al. / Journal of Sound and Vibration xxx (2018) 1e89
89
[189] P. Klein, J.F. Hamet, F. Anfosso-Ledee, An envelopment procedure for tyre/road contact, in: Proceedings of the 5th Symposium on Pavement Surface Characteristics, SURF, Toronto, Canada, 2004, 2004. [190] P. Klein, T. Beckenbauer, J.-F. Hamet, W. Kropp, Tyre/road noise prediction: a comparison between the SPERoN and HyRoNE models - Part 2, in: Proceedings of Acoustics08, Paris, France, 2008. [191] A. Kuijpers, B. Peeters, W. Kropp, T. Beckenbauer, Acoustic Optimization Tool e modeling refinements in the SPERoN framework, in: Technical Report MþP.DWW.06.04.7, MþP, Vught, 2007. [192] E. Rustighi, S.J. Elliott, S. Finnveden, K. Guly, T. Mocsai, M. Danti, Linear stochastic evaluation of tyre vibration due to tyre/road excitation, J. Sound. Vib. 310 (4e5) (2008) 1112e1127. [193] K.L. Johnson, Contact mechanics, J. Am. Chem. Soc. 37 (1985) 1e17. [194] J.D. Robson, Road surface description and vehicle response, Int. J. Veh. Des. 1 (1) (1979) 25e35. [195] G. Dubois, J. Cesbron, F. Anfosso-Ledee, H.P. Yin, A multi-asperity tyre/road contact model for rolling noise prediction, in: Forum Acusticum 2011, 2011. [196] G.D.J. Cesbron, H.P. Yin, F. Anfosso-Ledee, Macro-scale approach for rough frictionless multi-indentation on a viscoelastic half-space, Wear 272 (2011). [197] J. Cesbron, F. Anfosso-Ledee, D. Duhamel, H. Ping Yin, D. Le Houedec, Experimental study of tyre/road contact forces in rolling conditions for noise prediction, J. Sound. Vib. 320 (1e2) (2009) 125e144. [198] G. Dubois, J. Cesbron, H.P. Yin, F. Anfosso-Ledee, D. Duhamel, Statistical estimation of low frequency tyre/road noise from numerical contact forces, Appl. Acoust. 74 (9) (2013) 1085e1093. [199] G. Dubois, J. Cesbron, H.P. Yin, F. Anfosso-Ledee, Numerical evaluation of tyre/road contact pressures using a multi-asperity approach, I. J. Mech. Sci. 54 (2012). [200] P. Klein, J. Cesbron, A 3D envelopment procedure for tyre belt radiated noise level prediction, in: Proceedings of the INTER-NOISE 2016-45th International Congress and Exposition on Noise Control Engineering: towards a Quieter Future, 2016, pp. 2230e2241. [201] R.J. Pinnington, A particle-envelope surface model for roadetyre interaction, Int. J. Solids Struct. 49 (3e4) (Feb. 2012) 546e555. [202] R.J. Pinnington, Tyre-road contact using a particle-envelope surface model, J. Sound. Vib. 332 (26) (2013) 7055e7075. [203] R.J. Pinnington, Tyre contact forces on a particulate road surface, in: Proceedings of the INTER-NOISE 2016-45th International Congress and Exposition on Noise Control Engineering: towards a Quieter Future, 2016, pp. 2207e2218. [204] P. Cao, X. Yan, W. Xiao, L. Chen, A prediction model to coupling noise of tire tread patterns and road texture, in: Proceedings of the 8th International ́ Transportation Professionals - Logistics: the Emerging Frontiers of Transportation and Development in China, Conference of Chinese Logisticśand 2008, pp. 2332e2338. [205] J. Winroth, C. Hoever, W. Kropp, T. Beckenbauer, The contribution of air-pumping to tyre/road noise, in: Proceedings of AIA-DAGA 2013, 2013, pp. 1594e1597. [206] Z. Mohamed, X. Wang, A deterministic and statistical energy analysis of tyre cavity resonance noise, Mech. Syst. Signal Process. 70e71 (2016) 947e957. [207] A. Le Bot, Z. Bazari, P. Klein, J. Lelong, Statistical analysis of vibration in tyres, J. Sound. Vib. 392 (2017) 187e199. [208] B. Wang, D. Duhamel, On the design and optimization of acoustic network resonators for tire/road noise reduction, Appl. Acoust. 120 (2017) 75e84. [209] J.-P. Dalmont, C.J. Nederveen, N. Joly, Radiation impedance of tubes with different flanges: numerical and experimental investigations, J. Sound. Vib. 244 (3) (Jul. 2001) 505e534. ́ ́ of Road Traffic Noise, Department for Transport Welsh Office, 1988. ISBN 0 11 550847 3. [210] Department for Transport, Calculation [211] B.K.H. Leung, C.M. Mak, Is the CRTN method reliable and accurate for Traffic Noise prediction in Hong Kong? HKIE Trans. Hong Kong Inst. Eng. 15 (2) ́ ́ (2008) 17e23. [212] National Physical Laboratory (NPL), Technical Guides - Calculation of Road Traffic Noise 1988, 2005. http://resource.npl.co.uk/acoustics/techguides/ crtn/. [213] Department for Transport (UK), Design Manual for Roads and Bridges, Department for Transport, Highways Agency, UK, 2008. [214] P. Abbott, P. Nelson, Converting the UK Traffic Noise Index LA10,18h to EU Noise Indices for the Purposes of Noise Mapping, 2002. TRL Limited, Project ́ ́ Report PR/SE/451/02. [215] H.K. Lam, M.L. Tam, Reliability analysis of traffic noise estimates in Hong Kong, Transpn Res. D. Hong Kong 3 (4) (1998) 239e248. [216] W.M. To, C.W. Ip, C.K. Lam, T.H. Yau, A multiple regression model for urban traffic noise in Hong Kong, J. Acoust. Soc. Am. 112 (2) (2002) 551e556. [217] S.K. Tang, T. K.K., Estimating traffic noise for inclined roads with freely flowing traffic, Appl. Acoust. Hong Kong 65 (2004) 171e181. [218] N.P. Abbott PG, Converting the UK Traffic Noise Index LA10, 18 H to EU Noise Indices for the Purposes of Noise Mapping, TRL Ltd., UK, 2002. [219] D.C. O'Malley, V. King, E. Kenny L, Assessing methodologies for calculating road traffic noise levels in Ireland e converting CRTN indicators to the EU indicators (Lden, Lnight), Appl. Acoust. 70 (2009) 284e296. [220] E. Murphy, E.A. King, Scenario analysis and noise action planning: modelling the impact of mitigation measures on population exposure, Appl. Acoust. 72 (8) (2011) 487e494. [221] French standard, Bruit des infrastructures de transports terrestres (Noise from land transport infrastructure), 2001. XPS 31-133. ^te du 5 mai 1995 relatif au bruit des infrastructures routieres (Order of 5 May 1995 on road infrastructure noise), J. Off. 109 (1996) 7753. NOR: [222] Arre ENVP9540148A, du 10 mai 1995, Artic. 6 (Official J. 10 May 1995, Artic. 6). [223] Guide du bruit des transports terrestres, fascicule prevision des niveaux sonores (Guide to land transport noise, fascicule prediction of sound levels), CETUR, 1980. [224] S. Service d'e tudessur les transports, les routes etleursame nagements (Transportation, Roads and Lights, Road traffic noise prediction 1 e Calculating sound emissions from road traffic, 2009. [225] D. Dutilleux, Road traffic noise, in: Noise Mapping in the EU: Models and Procedures, CRC Press. Taylor Fr. Group, Boca Raton, FL, 2013, p. 77. tra, Road Noise Prediction - Part 2: Noise Propagation Computation Method Including Meteorological Effects (NMPB2008), 2009, pp. 1e136. [226] Se [227] Road Traffic Noise Reducing Devices. Test Method for Determining the Acoustic Performance. Normalized Traffic Noise Spectrum, 1998. BS EN 1793-3. [228] ISO, Acoustics e Attenuation of Sound during Propagation Outdoors e Part 1: Calculation of the Absorption of Sound by the Atmosphere, 1993. ISO 9613-1. [229] E. Murphy, E.A. King, Environmental Noise Pollution, 2014. [230] L. Dai, H.S. Lee, Ray acoustics approach to quantitative prediction of highway traffic noise, in: ASME International Mechanical Engineering Congress and Exposition, Proceedings (IMECE), vol. 8, 2010, pp. 223e228 no. PARTS A AND B. [231] ISO, Acoustics e Attenuation of Sound during Propagation Outdoors e Part 2: General Method of Calculation, 1996. ISO 9613-2. [232] D. Sheadel, Y. Ma, J. Cao, Meteorological effects on sound propagation, in: Spring Noise Conference on Environmental and Occupational Noise, Banff, 2007. [233] Y. Wei, Q. Feng, H. Wang, M. Kaliske, A hybrid numerical-experimental analysis for tire air-pumping noise with application to pattern optimization, Noise Control Eng. J. 64 (1) (2016) 56e63.
́
̀ ́
Please cite this article in press as: T. Li et al., Literature review of models on tire-pavement interaction noise, Journal of Sound and ́ https://doi.org/10.1016/j.jsv.2018.01.026 ́ Vibration (2018),