Next Article in Journal
Taking Advantage of Phosphate Functionalized Waterborne Acrylic Binders to Get Rid of Inhibitors in Direct-to-Metal Paints
Next Article in Special Issue
Review on the Relationship between Nano Modifications of Geopolymer Concrete and Their Structural Characteristics
Previous Article in Journal
Terephthalaldehyde–Phenolic Resins as a Solid-Phase Extraction System for the Recovery of Rare-Earth Elements
Previous Article in Special Issue
Polyamide 6-Aluminum Assembly Enhanced by Laser Microstructuring
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Determination of Vibroacoustic Parameters of Polyurethane Mats for Residential Building Purposes

by
Krzysztof Nering
* and
Alicja Kowalska-Koczwara
Faculty of Civil Engineering, Cracow University of Technology, 31-155 Cracow, Poland
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(2), 314; https://doi.org/10.3390/polym14020314
Submission received: 12 October 2021 / Revised: 13 December 2021 / Accepted: 17 December 2021 / Published: 13 January 2022
(This article belongs to the Special Issue Polymer Composites for Structural Applications)

Abstract

:
This paper is aimed at investigating the use of polyurethane mats, usually used as ballast mats, for residential building purposes. Ballast mats have features that may improve the vibroacoustic comfort in residential rooms. Their strength is certainly an advantage, along with vibration and acoustic insulation. However, the problem that an engineer has to deal with, for example in modeling these types of mats, is a limited knowledge of the material’s vibroacoustic parameters. Knowledge of these may be useful for residential buildings. This paper presents measurements of the vibroacoustic parameters of polyurethane mats, together with a suitable methodology and some results and analysis. The two main material parameters responsible for vibroacoustic protection were measured: the dynamic stiffness, which is related to the acoustic properties of the material, and the critical damping coefficient, which is obviously responsible for damping. The measurement methodology is clearly described. A total of five polyurethane materials with different densities were tested. It was possible to identify a relationship between the material density and the vibroacoustic parameters, which could offer an indication of which material to use, depending on the stimulus affecting a human in a given location.

1. Introduction

We live more and more in urbanized spaces where the emphasis is on quick movement from one place to another. Whether we like it or not, roads and railways must therefore be situated close to our living spaces. The proximity to infrastructure has its advantages and disadvantages. On the one hand is the proximity of work, shops or cultural centers, and on the other hand, noise and vibrations disturb our rest after a day’s work. Long-term exposure to noise and vibrations can not only be a nuisance but may also contribute to the deterioration of our health. Exposure to long-term or excessive noise can cause a range of health problems ranging from stress [1], poor concentration [2] and loss of productivity in the workplace [3,4] and communication difficulties and fatigue from lack of sleep [5], to more serious issues such as cardiovascular disease, cognitive impairment, tinnitus and hearing loss. The cardiovascular effects of long-term noise include an increase in blood pressure and heart rate [6,7]. Noise also has a negative effect on attention, working memory and episodic recall [8]. One of the worst conditions, of course, is hearing loss [9,10]. However, it is worth remembering that this happens very rarely and mainly applies to employees exposed to prolonged noise without the use of appropriate health and safety measures.
Most researchers in the context of transport impacts focus on noise as a factor that can be an annoyance and neglect the impact of vibrations, and especially the combined effect of vibrations and noise. Furthermore, low-frequency vibrations like transport vibrations are in the most dangerous range for our health. Low-frequency vibrations are vibrations in the 5–25 Hz range. They are dangerous because this frequency range is similar to the resonance frequency of the internal organs of our body [11,12]. Therefore, for vibrations in this range of frequencies, disturbing symptoms may appear, resulting from long-term exposure (see Table 1).
The combined effect of vibrations and noise may be more serious than when these stimuli are considered separately [13,14].
Therefore, one of the problems in our cities is pollution caused by noise and vibration. We want to live close to all the amenities that the city gives us, but we also want to live comfortably, without excessive noise or vibrations. We currently know three methods of reducing vibration and noise in our homes. The choice of method depends on the stage of construction of the vibration source and/or the building. When the source of vibrations (road, tram or railway) is under construction, the simplest and most common legal obligation [15,16] is to isolate it from the environment, for example by using sub-ballast mats. When the source already exists and the building is under construction, the building should be designed to meet all comfort requirements by using either vibro-insulating barriers in the ground [17] or a vibroacoustic floor [18,19]. The first of these two solutions is quite expensive, and the building owner must have the right to the land in the direction of propagation in order to build such a large structure. The second is much cheaper and requires only good selection of vibro-insulating materials. The problem lies in what material to choose and how to properly design the floor. Parameters such as the dynamic stiffness [20], related to protection against impact noise, or the critical damping ratio [21], related to protection against vibrations, may be helpful in selecting the appropriate vibroacoustic materials.
Mats used in railway or tramway construction as under-sleeper pads (USPs) are mostly tested under heavy load, due to their applications. Mainly, the three basic parameters are tested: static (Cstat) and dynamic (Cdyn) bedding moduli [22] and the loss tangent (tan φ) [23]. Static and dynamic bedding moduli are used for defining the dynamic stiffening ratio (stiffening coefficient) [24]. For residential building purposes, much better parameters to use are the dynamic stiffness ratio per unit area, which does not require the heavy loading needed for Cstat and Cdyn determination, and the critical damping ratio, instead of the loss tangent.
Mats used as USPs are tested according to their purpose. Usually, their purpose is to reduce the noise transmitted from the rails and the wheels of the railway wagon body to the environment, and which constitutes not only pollution but also annoyance. Therefore, the manufacturers of track mats provide noise-related parameters such as dynamic stiffness or impact sound pressure level. However, manufacturers also provide static and dynamic parameters related to the possible range of applications, such as the static modulus of elasticity, the dynamic modulus of elasticity, the coefficient of friction, etc. Very rarely, however, do manufacturers provide a parameter related to vibration damping, which is of key importance not only in reducing vibrations but also, through this reduction, reducing noise emissions. Getzner is practically the only manufacturer of track mats that makes data related to vibration damping publicly available. In its data sheets, the firm gives a parameter called the mechanical loss factor, which is, in fact, the loss tangent reduced to one value (not given over the whole frequency range). For example, the Sylomer SR 11 12.5 mm pad has a declared mechanical loss factor value of 0.25, while the Sylomer SR 110 12.5 mm pad has a mechanical loss factor of 0.14.

2. Methodology of Determining Vibroacoustic Parameters

Dynamic stiffness per unit area in the system can be obtained using Equation (1) acc. [25]:
s = F S Δ d ,
where S is the area of the test sample (m2), F is the dynamic force acting perpendicularly on the test sample (N) and Δd is the resultant dynamic change in the thickness of the elastic material (m).
The determination of the dynamic stiffness per unit area s’ of the test sample was performed by the resonance method: the resonance frequency fi of the basic vertical vibrations of the mass–spring–damper system is measured, where the test sample is the spring–damper of the tested system and the mass is the mass of the pressure plate.
The critical damping ratio (D) was obtained at the post-processing stage. Based on the acceleration frequency response spectrum, the logarithmic damping decrement (δi) was determined by the half-power method using Equation (2) [26,27] and the methodology presented in Figure 1.
Δ f i f i = δ i π 1 ( δ i 2 π ) 2 ,
where fi is the resonance frequency (Hz), Δfi is the frequency range in which the maximum value of the vibration acceleration at the resonant frequency (aR) is equal 0.707 (see Figure 1) and δi is the logarithmic damping decrement (–).
δ = 2 π D 1 D 2 ,
When Cstat and Cdyn are measured, a simple equation allows the dynamic stiffening ratio to be calculated [24]:
κ ( f ) = C d y n ( f ) C s t a t ,
There is also a Equation which describes the relationship between the loss tangent (tan φ) and the logarithmic damping decrement (δ) [28,29]:
tan φ = ( δ π ) [ 1 ( δ 2 π + ) ] ,
tan φ = δ π ,   f o r   δ < 2
Equation (5) can be simplified to the form of Equation (6), which applies to small values of the mechanical loss, which is equivalent to small values of damping. When considering damping in construction, we are generally dealing with very small values.

3. Measurement Methodology

A single-degree-of-freedom (SDOF) dynamic system is shown in Figure 2 acc. [30]), consisting of a mass–spring–damper system.
To recreate the schema presented in Table 2 in a real-world situation, the machine shown in Figure 3 was used to perform tests to obtain the vibroacoustic parameters. This machine is primarily used for testing dynamic stiffness [25]. It consists of a dynamic exciter which applies a harmonic force through a force sensor to the load plate, with a pre-loaded static force of 0.1–0.4 N. The load plate (8 kg) was placed over the tested sample using plaster of Paris in order to reduce any unevenness of the sample. The response of the system was measured using an IEPE (integrated electronics piezo-electric) accelerometer (mechanical sensors typically made of silicon, coupled with microelectronic circuits to measure the acceleration) with a sensitivity of 100 mV/g. The detailed characteristics of the devices used in the dynamic stiffness machine are presented in Table 2.
The materials chosen for the measurements were flexible rebound polyurethane foams (Type R) with varying densities from a nominal 150 kg/m3 to 250 kg/m3 (150, 160, 180, 200 and 250 kg/m3). The choice was dictated by the fact that in this density range, the foam is produced by the same technological process. For each nominal density, 3 samples were tested. The actual density of samples is presented in Table 3.
The dimensions of the samples for measurement were 200 mm × 200 mm × 50 mm. This is due to the fact that machine used in this test procedure was adapted to using 200 mm × 200 mm samples. The thickness can vary, in terms of the requirements of the testing machine. Considering the resonant frequency, the thickness must be high enough to provide a resonant frequency in range of 1 Hz to 100 Hz. An example of a tested sample is shown in Figure 4. It is rebound polyurethane foam with a nominal density of 160 kg/m3. The actual density of the sample in the picture was 156.5 kg/m3.
Dynamic excitation of the testing system consisted of a force generated by the exciter modulated by a sinusoidal signal. The amplitude of the applied sinusoidal force was 0.2 N ± 0.005 N and the frequency range was from 1Hz to 100 Hz. The frequency was increased every 1 s of the measurement time by 0.1 Hz. The system response was measured using an IEPE accelerometer located at the load plate. An example of the excitation force and system response at 100 Hz is presented in Figure 5.
The results of the measured accelerations and the force presented in Figure 5 are the values measured directly by the relevant sensors. Apart from the measured frequency at a given moment, the signal also contains the self-noise of the exciter and other dynamic disturbances coming from the environment, which are not completely removed by vibro-isolation of the base plate. A Butterworth low-pass filter was used before the signal was converted to its resultant amplitude to reduce possible contamination of the acceleration values used for the analysis. Possible phase shifts due to the use of the filter were ignored as only the signal amplitude is taken into account in the results.
According to Equations (2) and (3), the critical damping ratio was calculated using the methodology illustrated in Figure 1. An example of this calculation for a sample of nominal density 160 kg/m3 is shown in Figure 6. The critical damping ratio for this example was D = 7.3% (Δfi was estimated from Figure 6 as 6.1 Hz).
The dynamic stiffness values were calculated according to Equation (1). In order to calculate the aforementioned values, the resonant frequency was approximated using the maximum value of the response spectrum. The indication of the resonant frequency is shown in Figure 7.

4. Measurement Results

4.1. Results for Rebound Polyurethane

According to the methodology described above, the two materials parameters (dynamic stiffness and critical damping ratio) were calculated for each rebound polyurethane foam sample. The results for these two parameters are listed in Table 4.
The results are also presented in the figures. In Figure 8, the dynamic stiffness results are shown and a trend line with the scattering of the results is presented.
The tested sample densities varied between 143.0 kg/m3 and 264.5 kg/m3. They fitted reasonably well to the extreme values of the nominal density of 150–250 kg/m3.
As can be seen from Table 4 and Figure 8, the measured dynamic stiffness was between 11 MN/m3 and 40 MN/m3. The value of R2 (0.978) is very high, and the root mean square error RMSE of 1.498 is relatively low. Therefore, the relationship between the dynamic stiffness and the density ρ for this type of material can be given as follows:
s = 0.2269   ρ 21.1
where s’ is the dynamic stiffness [MN/m3] and ρ is the density [kg/m3].
The critical damping ratio can be presented in the same way. In Figure 9, the results for the damping ratio are shown, with a trend line suitable for the relationship between damping and density.
The critical damping ratio values were between 6.87% and 8.54%. The value of R2 (0.9122) is high, and the root mean square error (RMSE) of 0.001527 is relatively low. The relationship between the critical damping ratio and the density of this type of material can be shown as follows:
D = 0.0001188   ρ   + 0.05466 ,
where D is the critical damping ratio (-) and ρ is the density (kg/m3).
Based on the fact that two separate quantities—dynamic stiffness and critical damping ratio—can be predicted by one parameter, i.e., the density, the dependency between dynamic stiffness and critical damping ratio is analyzed below in Figure 10.
The fitting is as good as in the previous Equations. The value of R2 (0.9483) is high and the root mean square error (RMSE) of 0.001172 is low. Hence, the Equation describing the relationship between the critical damping ratio and the dynamic stiffness for rebound polyurethane foam, with densities varying from a nominal 150 kg/m3 to 250 kg/m3, can be written as follows:
D = 0.0004968   s   + 0.06497
where D is the critical damping ratio (-) and s’ is the dynamic stiffness (MN/m3).

4.2. Control Samples

In order to evaluate the correctness of the methodology used for measurement and calculation of the dynamic stiffness and critical damping ratio, control samples of known products were used. Due to fact that manufacturers in general do not test their products for both dynamic stiffness and damping, two different materials had to be chosen as control samples. For the dynamic stiffness measurements, Ursa TEP mineral wool was tested. For the damping, the elastomer polyurethane Sylomer SR 11 from Getzner was selected as the control sample.
Considering mineral wool, it is necessary to underline the fact that the tested mineral wool has a relatively low flow resistivity (~20 kPas/m2). This means that to measure the dynamic stiffness, the dynamic stiffness of the enclosed gas must be added. The dynamic stiffness of the enclosed gas s’a is given by the Equation:
s a = p 0 d ε ,
where p0 is the atmospheric pressure during the test (1023 hPa), d is the sample thickness and ε is the porosity of the test specimen.
The results for the Ursa TEP dynamic stiffness measurements are presented in Table 5.
An example resonance curve is presented in Figure 11.
For the Sylomer damping test, it was important to recalculate the critical damping factor D’s mechanical loss factor (estimated using Equation (5b)) to a quantity given by the manufacturer. This quantity was compared with the manufacturer’s data.
As can be seen in Table 6, the results do not agree exactly with manufacturer’s data but they are within an acceptable range of error. An example frequency graph for the Sylomer sample is shown in Figure 12.

5. Practical Applications

The dynamic stiffness is used to evaluate the ability of a floating floor system to reduce impact sounds in dwellings. The parameter describing the damping in such a floor is responsible for the damping of vibrations that can come both from internal sources (e.g., a machine working above the ceiling) or from external sources propagating through the ground into the building, such as traffic. Examples of floor structures with elements damping both impact sounds and vibrations are shown in Figure 13 and Figure 14. Floating floors have become increasingly popular for many types of floor coverings. The term “floating floor” does not refer to a type of flooring material, but rather to a method of installation that can be used with a variety of materials, including laminates, engineered hardwood and luxury vinyl flooring. In this method, individual planks (or in some cases tiles) interlock edge-to-edge to form a single mat-like surface that simply rests on the underlayment. It is quite different from the glue-down or nail-down methods that are still used for ceramic and stone tiles, and which were once standard for nearly all flooring materials.
Based on the dynamic stiffness results, the analysis covered the possibility of reducing impact sounds depending on the density/damping of the polyurethane used and the density of the entire floating floor system. In Figure 15, the possibility of reducing the weighted impact sound pressure level of floating floor system screeds made of sand/cement or calcium sulfate with rebound polyurethane foam as a resilient layer is shown.
The same analysis was performed for the second type of floating floor (asphalt or dry), as shown in Figure 16.
In order to provide relatively smooth graphs for the results presented above, averaging of results using nominal density was performed. The density and dynamic stiffness values were linearly averaged in each group of nominal densities. After obtaining the average density and the corresponding dynamic stiffness for each nominal density, the method in Annex C of [32] was applied to obtain the reduction in impact sound level for varying thicknesses. To create the iso-surfaces presented in Figure 15 and Figure 16, spline smoothing was performed between the points obtained using the above method. The fact that decreasing the damping ratio leads to a lower reduction of impact noise may be counterintuitive. It should be underlined that the ability to reduce impact noise in terms of building acoustics depends mainly on the dynamic stiffness of the resilient layer. The damping ratio is correlated with the density (see Euqation (8)) and the dynamic stiffness (see Equation (9)). However, it should be remembered that the correlations presented in this paper are for rebound polyurethane and applying these to different materials may be misleading. The strict influence of damping on the reduction of impact noise is not the topic of this paper.
It is worth emphasizing that both floating floor systems meet the requirements for impact sound insulation, the maximum level of which in Poland is 55 dB in multifamily houses, considering high-mass floating floors and low-density polyurethanes.

6. Discussion and Conclusions

This article presents a measurement methodology for the vibroacoustic parameters for rebound polyurethane foam, which could be used in floating floor systems to reduce the impact sound level and to increase the damping ratio coefficient. Although floating floor systems are widely used to reduce noise, they are not used to reduce vibration. However, as shown in the article, this type of flooring system can have vibration-reducing properties, for vibrations from both from external and internal sources, due to polyurethane’s damping properties.
Rebound polyurethane foam has comparable or even better parameters of weighted reduction impact sound pressure level than the mineral wool or elasticized Styrofoam commonly used in civil engineering [33,34]. An example comparison of the impact sound reduction ranges is given in Table 7. This leads to the conclusion that rebound polyurethane is a worthy alternative to commonly used materials as a resilient layer for floating floor systems.
It is worth noting that higher densities of rebound polyurethane tend to have relatively high damping parameters. The critical damping ratio at a level of 8% is useful when a building is located in dynamic influence zones. This value of damping indicates that rebound polyurethane can be used not only as acoustic insulation but also as a material responsible for the reduction of the vibration perception of the building’s residents [35]. Of course, it must be remembered that an increase in damping increases the dynamic stiffness of polyurethane, which leads to a decrease (sometimes acceptable) in the acoustic performance of the floating floor.
This paper presents methods for the prediction of vibroacoustic parameters of specific rebound polyurethane samples based only on their density. This leads to simplification and acceleration of the design and execution process for dwellings using this type of material.
Damping, which is helpful for vibration reduction in dynamic influence zones, is inversely proportional to the noise reduction capabilities. This leads to the conclusion that during the design process, an analysis must be conducted to determine the “sweet spot” where acoustical advantages will not overcome the vibration aspect. A method to address this problem can be found in [36].

Author Contributions

Conceptualization, K.N. and A.K.-K.; methodology K.N. and A.K.-K.; software, K.N.; validation, K.N.; formal analysis, K.N. and A.K.-K.; investigation, K.N. and A.K.-K.; resources, K.N. and A.K.-K.; data curation, K.N.; writing—original draft preparation, K.N. and A.K.-K.; writing—review and editing, K.N. and A.K.-K.; visualization, K.N.; supervision, A.K.-K.; project administration, K.N. and A.K.-K.; funding acquisition, K.N. and A.K.-K. All authors have read and agreed to the published version of the manuscript.

Funding

Scientific research results were financed by the European Union from the European Regional Development Fund within the Smart Growth Operational Programme 2014-2020. “The innovative technology of vibro-acoustic floor insulation” project is implemented as a part of the Regional Science and Research Agendas (RANB) competition of the National Centre for Research and Development (NCRD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Babisch, W. Stress hormones in the research on cardiovascular effects of noise. Noise Health 2003, 5, 1–11. [Google Scholar]
  2. Björk, J.; Ardö, J.; Stroh, E.; Lövkvist, H.; Östergren, P.-O.; Albin, M. Road traffic noise in southern Sweden and its relation to annoyance, disturbance of daily activities and health. Scand. J. Work. Environ. Health 2006, 32, 392–401. [Google Scholar] [CrossRef] [Green Version]
  3. Aybeka, A.; Kamer, H.A.; Arslan, S. Personal noise exposures of operators of agricultural tractors. Appl. Ergon. 2010, 41, 274–281. [Google Scholar] [CrossRef]
  4. Rubio-Romero, J.C.; Carrillo-Castrillo, J.A.; Soriano-Serrano, M.; Galindo-Reyes, F.; de la Varga-Salto, J. A longitudinal study of noise exposure and its effects on the hearing of olive oil mill workers. Int. J. Ind. Ergon. 2018, 67, 60–66. [Google Scholar] [CrossRef]
  5. Halperin, D. Environmental noise and sleep disturbances: A threat to health? Sleep Sci. 2014, 7, 209–212. [Google Scholar] [CrossRef] [Green Version]
  6. Münzel, T.; Gori, T.; Babisch, W.; Basner, M. Cardiovascular effects of environmental noise exposure. Eur. Heart J. 2014, 35, 829–836. [Google Scholar] [CrossRef] [Green Version]
  7. Münzel, T.; Sørensen, M.; Daiber, A. Transportation noise pollution and cardiovascular disease. Nat. Rev. Cardiol. 2021, 18, 619–636. [Google Scholar] [CrossRef]
  8. Wright, B.; Peters, E.; Ettinger, U.; Kuipers, E.; Kumari, V. Understanding noise stress-induced cognitive impairment in healthy adults and its implications for schizophrenia. Noise Health 2014, 16, 166–176. [Google Scholar] [CrossRef]
  9. Ding, T.; Yan, A.; Liu, K. What is noise-induced hearing loss? Br. J. Hosp. Med. 2019, 80, 525–529. [Google Scholar] [CrossRef] [Green Version]
  10. Rodrigues, H.F.S.; Filho, F.J.M.B.D.O.; Ferraz, D.P.; Neto, A.F.D.A.; Torres, S.; Metidieri, M.M. Noise-Induced Hearing Loss (NIHL): Literature review with a focus on occupational medicine. Int. Arch. Otorhinolaryngol. 2014, 17, 208–212. [Google Scholar] [CrossRef] [Green Version]
  11. .Coermann, R.R. The Mechanical Impedance of the Human Body in Sitting and Standing Position at Low Frequencies. Hum. Factors J. Hum. Factors Ergon. Soc. Oct. 1962, 4, 227–253. [Google Scholar] [CrossRef]
  12. Pradko, F.; Lee, R.; Kaluza, V. Theory of Human Vibration Response. S&T Reports 1966. Available online: https://apps.dtic.mil/sti/pdfs/AD0634632.pdf (accessed on 1 October 2021).
  13. Lee, P.J.; Griffin, M.J. Combined effect of noise and vibration produced by high-speed trains on annoyance in buildings. J. Acoust. Soc. Am. 2013, 133, 2126–2135. [Google Scholar] [PubMed] [Green Version]
  14. Peris, E.; Woodcock, J.S.; Sica, G.; Moorhouse, A.T.; Waddington, D.C. Annoyance from railway vibration in residential environments: Factors of importance when considering exposure-response relationships. J. Acoust. Soc. Am. 2012, 131, 3504. [Google Scholar] [CrossRef] [Green Version]
  15. The Act of April 27, 2001, Environmental Protection Law. Journal of Laws 2001 No.62, Item. 627. Available online: https://www.global-regulation.com/translation/poland/10093814/the-act-of-27-april-2001%252c-the-environmental-protection-law.html (accessed on 1 October 2021).
  16. Regulation of the Minister of Infrastructure and Construction of November 14, 2017 Amending the Regulation on Technical Conditions to be Met by Buildings and Their Location. Available online: https://www.global-regulation.com/translation/poland/3353940/regulation-of-the-minister-of-infrastructure-of-12-april-2002-on-technical-conditions%252c-which-should-correspond-to-the-buildings-and-their-location.html (accessed on 1 October 2021).
  17. Andersen, L.; Nielsen, S. Reduction of ground vibration by means of barriers or soil improvement along a railway track. Soil Dyn. Earthq. Eng. 2005, 25, 701–716. [Google Scholar] [CrossRef]
  18. Hasheminejad, S.M.; Vesal, R. Numerical simulation of impact sound transmission control across a smart hybrid double floor system equipped with a genetically-optimized NES absorber. Appl. Acoust. 2021, 182, 108179. [Google Scholar] [CrossRef]
  19. Kłosak, A.; Kowalska-Koczwara, A.; Pachla, F.; Stypuła, K.; Tatara, T.; Zając, B. Proposal of new vibro-acoustic floor. MATEC Web Conf. 2018, 211, 10001. [Google Scholar] [CrossRef]
  20. Wu, K.; Kuhlenkoetter, B. Experimental Analysis of the Dynamic Stiffness in Industrial Robots. Appl. Sci. 2020, 10, 8332. [Google Scholar] [CrossRef]
  21. Inman, D. Critical damping. In Encyclopedia of Vibration; Braun, S., Ed.; Elsevier: Amsterdam, The Netherlands, 2001. [Google Scholar]
  22. Zbiciak, A.; Kraśkiewicz, C.; Al Sabouni-Zawadzka, A.; Pełczyński, J.; Dudziak, S. A Novel Approach to the Analysis of Under Sleeper Pads (USP) Applied in the Ballasted Track Structures. Materials 2020, 13, 2438. [Google Scholar] [CrossRef] [PubMed]
  23. Gandel’sman, M.I.; Gotlib, Y.Y.; Darinskii, A.A. Frequency dependence of the mechanical loss tangent for a system of two-block polymer chains. Polym. Sci. USSR 1981, 23, 2349–2358. [Google Scholar] [CrossRef]
  24. Kraśkiewicz, C.; Zbiciak, A.; Oleksiewicz, W.; Karwowski, W. Static and dynamic parameters of railway tracks retrofitted with under sleeper pads. Arch. Civ. Eng. 2018, 64, 187–201. [Google Scholar] [CrossRef] [Green Version]
  25. EN 29052-1:2011/ISO 9052-1. Acoustics—Determination of Dynamic Stiffness—Part 1: Materials Used under Floating Floors in Dwellings. Available online: https://standards.iteh.ai/catalog/standards/cen/04a2dad0-ff01-4409-b2d4-6d6d62feb5c7/en-29052-1-1992 (accessed on 1 October 2021).
  26. Papagiannopoulos, G.A.; Hatzigeorgiou, G.D. On the use of the half-power bandwidth method to estimate damping in building structures. Soil Dyn. Earthq. Eng. 2011, 31, 1075–1079. [Google Scholar] [CrossRef]
  27. Wu, B. A correction of the half-power bandwidth method for estimating damping. Arch. Appl. Mech. 2014, 85, 315–320. [Google Scholar] [CrossRef]
  28. Magalas, L.B.; Malinowski, T. Measurement Techniques of the Logarithmic Decrement. Solid State Phenom. 2003, 89, 247–260. [Google Scholar] [CrossRef]
  29. Kwiatkowski, D. Assessment of dynamic properties of the composite of PA/PP mixture with glass fiber. Composites 2003, 3, 322–324. (In Polish) [Google Scholar]
  30. Irvine, T. An Introduction to Shock & Vibration Response Spectra. 2019. Available online: https://www.scribd.com/document/407762633/ebook-tom-irvine-shock-vibration-response-spectra-pdf (accessed on 1 October 2021).
  31. Buryachenko, V.; Iarve, E.; Kim, R.; Sihn, S.; Tandon, G.P. Aerospace Composite Materials: Delivery Order 0002: Development and Validation of Micromechanical Models for Composites. 2002, 167. Available online: https://www.researchgate.net/publication/235020753_Aerospace_Composite_Materials_Delivery_Order_0002_Development_and_Validation_of_Micromechanical_Models_for_Composites (accessed on 1 October 2021).
  32. PN-EN ISO 12354-2:2017-10. Building Acoustics—Determination of Acoustic Properties of Buildings on the Basis of Component Properties—Part 2: Impact Sound Insulation between Rooms. Available online: https://www.iso.org/standard/70243.html (accessed on 1 October 2021).
  33. Miškinis, K.; Dikavičius, V.; Ramanauskas, J.; Norvaišienė, R. Dependence between Reduction of Weighted Impact Sound Pressure Level and Specimen Size of Floating Floor Construction. Mater. Sci. 2012, 18, 93–97. [Google Scholar] [CrossRef] [Green Version]
  34. Nowotny, Ł.; Nurzynski, J. The Influence of Insulating Layers on the Acoustic Performance of Lightweight Frame Floors Intended for Use in Residential Buildings. Energies 2020, 13, 1217. [Google Scholar] [CrossRef] [Green Version]
  35. Nelson, P. Vibration isolation on floating floors. Appl. Acoust. 1982, 15, 97–109. [Google Scholar] [CrossRef]
  36. Nering, K.; Kowalska-Koczwara, A.; Stypuła, K. Annoyance Based Vibro-Acoustic Comfort Evaluation of as Summation of Stimuli Annoyance in the Context of Human Exposure to Noise and Vibration in Buildings. Sustainability 2020, 12, 9876. [Google Scholar] [CrossRef]
Figure 1. Half-power method illustration [26,27]. In order to obtain the critical damping ratio (D) an appropriate algebraic operation should be applied to the measured value of the logarithmic damping decrement (see Equation (3)) acc. [28].
Figure 1. Half-power method illustration [26,27]. In order to obtain the critical damping ratio (D) an appropriate algebraic operation should be applied to the measured value of the logarithmic damping decrement (see Equation (3)) acc. [28].
Polymers 14 00314 g001
Figure 2. Single-degree-of-freedom mass–spring–damper system [30], where m is the mass of the system, c is the damping, k is the stiffness, f is the excitation force and x is the displacement.
Figure 2. Single-degree-of-freedom mass–spring–damper system [30], where m is the mass of the system, c is the damping, k is the stiffness, f is the excitation force and x is the displacement.
Polymers 14 00314 g002
Figure 3. Dynamic stiffness testing machine for testing mass–spring–damper systems: (a) axonometry view; (b) front view. One tile on tile scale is 10 mm × 10 mm (own elaboration).
Figure 3. Dynamic stiffness testing machine for testing mass–spring–damper systems: (a) axonometry view; (b) front view. One tile on tile scale is 10 mm × 10 mm (own elaboration).
Polymers 14 00314 g003
Figure 4. An example of tested samples of rebound polyurethane foam: (a) top view; (b) side view. One tile on the tile scale is 10 mm × 10 mm (own elaboration).
Figure 4. An example of tested samples of rebound polyurethane foam: (a) top view; (b) side view. One tile on the tile scale is 10 mm × 10 mm (own elaboration).
Polymers 14 00314 g004
Figure 5. Force applied to tested system and its acceleration response at 100 Hz: (a) pure measured signal; (b) filtered signal for analysis (own elaboration).
Figure 5. Force applied to tested system and its acceleration response at 100 Hz: (a) pure measured signal; (b) filtered signal for analysis (own elaboration).
Polymers 14 00314 g005
Figure 6. Acceleration frequency response spectrum of an example sample with indication of half-power frequency range (own elaboration).
Figure 6. Acceleration frequency response spectrum of an example sample with indication of half-power frequency range (own elaboration).
Polymers 14 00314 g006
Figure 7. Acceleration frequency response spectrum of an example sample with indication of approximate resonant frequency (own elaboration).
Figure 7. Acceleration frequency response spectrum of an example sample with indication of approximate resonant frequency (own elaboration).
Polymers 14 00314 g007
Figure 8. Results for dynamic stiffness depending on density of material, with fitted curve and 95% confidence bounds (own elaboration).
Figure 8. Results for dynamic stiffness depending on density of material, with fitted curve and 95% confidence bounds (own elaboration).
Polymers 14 00314 g008
Figure 9. Results for critical damping ratio depending on density of material, with 95% confidence bounds (own elaboration).
Figure 9. Results for critical damping ratio depending on density of material, with 95% confidence bounds (own elaboration).
Polymers 14 00314 g009
Figure 10. Results for critical damping ratio depending on dynamic stiffness of material, with 95% confidence bounds (own elaboration).
Figure 10. Results for critical damping ratio depending on dynamic stiffness of material, with 95% confidence bounds (own elaboration).
Polymers 14 00314 g010
Figure 11. Acceleration frequency response spectrum of an example sample of Ursa TEP 23 mm mineral wool with indication of resonant frequency (own elaboration).
Figure 11. Acceleration frequency response spectrum of an example sample of Ursa TEP 23 mm mineral wool with indication of resonant frequency (own elaboration).
Polymers 14 00314 g011
Figure 12. Acceleration frequency response spectrum of an example sample of Sylomer SR 11 12.5 mm polyurethane with indication of half-power frequency range (D = 9%) (own elaboration).
Figure 12. Acceleration frequency response spectrum of an example sample of Sylomer SR 11 12.5 mm polyurethane with indication of half-power frequency range (D = 9%) (own elaboration).
Polymers 14 00314 g012
Figure 13. An example of floating floor system screeds made of sand/cement or calcium sulfate with rebound polyurethane foam as a resilient layer (own elaboration).
Figure 13. An example of floating floor system screeds made of sand/cement or calcium sulfate with rebound polyurethane foam as a resilient layer (own elaboration).
Polymers 14 00314 g013
Figure 14. An example of asphalt floating floor or dry floating floor constructions with rebound polyurethane foam as a resilient layer (own elaboration).
Figure 14. An example of asphalt floating floor or dry floating floor constructions with rebound polyurethane foam as a resilient layer (own elaboration).
Polymers 14 00314 g014
Figure 15. Weighted reduction of impact sound pressure level for floating floor screeds made of sand/cement or calcium sulfate with rebound polyurethane foam as a resilient layer, with varying densities and critical damping ratios (own elaboration).
Figure 15. Weighted reduction of impact sound pressure level for floating floor screeds made of sand/cement or calcium sulfate with rebound polyurethane foam as a resilient layer, with varying densities and critical damping ratios (own elaboration).
Polymers 14 00314 g015
Figure 16. Weighted reduction of impact sound pressure level for asphalt floating floor or dry floating floor constructions with rebound polyurethane foam as a resilient layer, with varying densities and critical damping ratios (own elaboration).
Figure 16. Weighted reduction of impact sound pressure level for asphalt floating floor or dry floating floor constructions with rebound polyurethane foam as a resilient layer, with varying densities and critical damping ratios (own elaboration).
Polymers 14 00314 g016
Table 1. Symptoms caused by vibration [11].
Table 1. Symptoms caused by vibration [11].
Symptomsf (Hz) 1
General feeling of discomfort4–9
Head symptoms13–20
Lower jaw symptoms6–8
Influence on speech13–20
“Lump in throat”12–16
Chest pains5–7
Abdominal pains4–10
Urge to urinate10–18
Increased muscle tone13–20
Influence on breathing movement4–8
Muscle constractions4–9
1f—frequency.
Table 2. Detailed characteristics of key components of dynamic stiffness machine.
Table 2. Detailed characteristics of key components of dynamic stiffness machine.
Device Name/ManufacturerKey FeatureKey Value of Parameters
Dynamic exciter—Brüel & Kjær Mini-shaker Type 4810Provides sinusoidal forceSine peak max 10 N
Frequency range DC-18 kHz
IEPE accelerometer—MMF KS78B.100Measures acceleration of system responsePeak acceleration 60 g (~600 m/s2)
Linear frequency range (5% deviation)
0.6 Hz–14 kHz
Force sensor—Forsentek FSSM 50 NMeasures force applied to systemCapacity 50N
Rated output 2.0mV/V
Hysteresis ± 0.1% R.O. (rated output)
Dynamic stiffness test benchMeasures resonant frequency of sample (200 mm × 200 mm) under load of 8 kgLinear frequency range upper limit
(5% deviation)
250 Hz—measured
Table 3. Nominal and actual density of tested samples, with Young’s modulus (own elaboration) and Poisson’s ratio [31].
Table 3. Nominal and actual density of tested samples, with Young’s modulus (own elaboration) and Poisson’s ratio [31].
Nominal Density (kg/m3)Sample IDActual Density (kg/m3)Young’s Modulus (MPa)Poisson’s Ratio (-)
[31]
25001261.02.10.23
02264.52.0
03253.51.6
20011215.01.50.23
12192.01.7
13209.01.3
18021165.01.00.23
22176.50.9
23174.00.9
16032161.00.70.24
32161.00.7
33151.50.6
15041150.50.50.24
42149.50.6
43143.00.5
Table 4. Results of measurement of dynamic stiffness and critical damping ratio for tested samples of rebound polyurethane foam.
Table 4. Results of measurement of dynamic stiffness and critical damping ratio for tested samples of rebound polyurethane foam.
Sample IDActual Density (kg/m3)Dynamic Stiffness (MN/m3)Critical Damping Ratio (-)
01261.0380.085
02264.5400.083
03253.5330.081
11215.0300.078
12192.0240.077
13209.0260.079
21165.0180.074
22176.5200.076
23174.0190.076
31161.0140.072
32161.0150.073
33151.5130.070
41150.5120.074
42149.5120.069
43143.0110.069
Table 5. Results of measurement of dynamic stiffness of tested samples of Ursa TEP 23 mm mineral wool, with CI = 95% (10 samples).
Table 5. Results of measurement of dynamic stiffness of tested samples of Ursa TEP 23 mm mineral wool, with CI = 95% (10 samples).
Material NameActual Density (kg/m3)Dynamic Stiffness (MN/m3)Dynamic Stiffness of Enclosed Gas (MN/m3)Total Dynamic Stiffness
Average
(MN/m3)
Declared Value of Dynamic Stiffness by Manufacturer (MN/m3)Difference
(MN/m3)
Ursa TEP 23 mm81.2 (±3.1)5.5 (±0.4)4.6 (±0.2)10.1110.9
Table 6. Results of measurement of dynamic stiffness of tested samples of Sylomer SR 11 12.5 mm (17 samples).
Table 6. Results of measurement of dynamic stiffness of tested samples of Sylomer SR 11 12.5 mm (17 samples).
Material NameActual Density (kg/m3)Critical Damping Factor (-)Mechanical Loss Factor (-)Declared Value of Mechanical Loss Factor by Manufacturer (-)Difference
(-)
Sylomer SR 11463.4 (±23.8)0.096 (±0.037)0.178
(±0.017)
0.250.072
Table 7. Comparison of impact sound pressure level reduction of rebound polyurethane with other materials. It is assumed that the floating slab is screed made of sand/cement or calcium sulfate with a surface density of 80 kg/m2 (according to the producers’ brochures).
Table 7. Comparison of impact sound pressure level reduction of rebound polyurethane with other materials. It is assumed that the floating slab is screed made of sand/cement or calcium sulfate with a surface density of 80 kg/m2 (according to the producers’ brochures).
MaterialRange of Impact Sound Pressure Level Reduction ΔLw (dB)
Tested rebound polyurethane23–31
Mineral wool24–34
Elasticized Styrofoam23–29
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nering, K.; Kowalska-Koczwara, A. Determination of Vibroacoustic Parameters of Polyurethane Mats for Residential Building Purposes. Polymers 2022, 14, 314. https://doi.org/10.3390/polym14020314

AMA Style

Nering K, Kowalska-Koczwara A. Determination of Vibroacoustic Parameters of Polyurethane Mats for Residential Building Purposes. Polymers. 2022; 14(2):314. https://doi.org/10.3390/polym14020314

Chicago/Turabian Style

Nering, Krzysztof, and Alicja Kowalska-Koczwara. 2022. "Determination of Vibroacoustic Parameters of Polyurethane Mats for Residential Building Purposes" Polymers 14, no. 2: 314. https://doi.org/10.3390/polym14020314

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

Article Metrics

Back to TopTop