• No results found

Real-time observation of waviness formation during C/PEEK consolidation

N/A
N/A
Protected

Academic year: 2021

Share "Real-time observation of waviness formation during C/PEEK consolidation"

Copied!
8
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Contents lists available atScienceDirect

Composites Part A

journal homepage:www.elsevier.com/locate/compositesa

Real-time observation of waviness formation during C/PEEK consolidation

E.T.M. Krämer

a,b

, W.J.B. Grouve

b

, S. Koussios

a

, L.L. Warnet

b

, R. Akkerman

a,b,⁎

aThermoPlastic composites Research Center (TPRC), Palatijn 15, 7521PN Enschede, Netherlands bUniversity of Twente, Drienerlolaan 5, 7522NB Enschede, Netherlands

A R T I C L E I N F O Keywords: Thermoplastic composites Waviness Consolidation Defects A B S T R A C T

Defects such as fiber waviness can occur during the production of thermoplastic composite parts. It is desirable to prevent or minimize fiber waviness as it causes a knockdown in the part’s mechanical properties. However, the mechanisms governing the formation of fiber waviness are currently not fully understood. A setup for the real-time observation of waviness formation during flat laminate consolidation is presented in this paper. For C/ PEEK unidirectional composites, it was observed that waviness forms during cooling between the maximum process temperature and the crystallization temperature. A higher processing temperature leads to waviness with a larger misalignment angle. No such trends were observed for the wavelength. Waviness does not form or increase in severity after crystallization. The driving force for waviness formation in flat laminate consolidation was determined to be tool shrinkage transferred to the laminate via tool-ply friction.

1. Introduction

1.1. Background

Thermoplastic composites (TPC) are increasingly used in the aero-space industry because of their excellent specific stiffness and strength, as well as their potential for cost-effective and automated manu-facturing. However, there are still some challenges in manumanu-facturing. A prominent one is the occurrence of defects such as fiber waviness during part production. Two examples of such parts are shown inFig. 1. Fiber waviness causes a knockdown in mechanical properties that de-pends on the shape of the wave and the number of affected plies[1–6]. Fiber waviness in thermoplastic composite parts can originate from the manufacturing stage, even when the part has been made following the design and manufacturing specifications. Components with such nonconformities need expensive additional non-destructive inspection (NDI). These parts are generally scrapped, as the influence of fiber waviness on the mechanical performance is poorly understood. The scrap rate can be reduced by minimizing or even preventing fiber wa-viness, which requires a proper understanding of the governing me-chanisms.

A flat laminate consolidated using a heated press or an autoclave may already suffer from fiber waviness, even in the absence of com-pressive fiber loads due to forming of the laminate. The formation of in-plane waviness during consolidation of a flat laminate is investigated in this paper, with a particular focus on the description of the governing

mechanisms. A short literature review is presented before setting the objectives.

1.2. Literature review

The imperfection of fiber straightness has been studied extensively in the past, with multiple designations in the literature: waviness, misalignment, marcelling, chicken tracks and wrinkles. The ground-work for the effects of fiber misalignment on the mechanical properties in thermoset composites (TSC) dates back as early as 1965[1]. The significance on the knockdown of part properties in high-performance TSCs has been established for out-of-plane waviness[3]and in-plane waviness[4]. Generally, waviness is considered unacceptable due to the large negative influence on part performance and uncertainties in ac-curately determining the knockdown[3]. Unfortunately, the amount of literature on the formation of waviness is less abundant. Moreover, a very limited amount of literature is available when focussing on TPCs. The available entries focus mainly on the forming process.

The mechanisms for the formation of fiber waviness in the literature can be divided into two loading conditions: axial compressive fiber stresses and transverse fiber stresses. Not all mechanisms have been investigated with respect to TPCs. However, a lot can be learned from research into their thermosetting counterparts.

1.2.1. Axial compressive fiber stress

The most reported loading condition for the formation of fiber

https://doi.org/10.1016/j.compositesa.2020.105872

Received 14 January 2020; Received in revised form 28 February 2020; Accepted 10 March 2020 ⁎Corresponding author at: University of Twente, Drienerlolaan 5, 7522NB Enschede, Netherlands.

E-mail address:r.akkerman@utwente.nl(R. Akkerman).

Composites Part A 133 (2020) 105872

Available online 17 March 2020

1359-835X/ © 2020 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).

(2)

waviness is an axial compressive force on the fiber reinforcement. Fibers may buckle when subjected to such forces, causing waviness in the ply[7]. Three scales of buckling can be distinguished: bucking on the macroscale (laminate), mesoscale (ply/tow) and microscale (fiber). Macroscale buckling may arise during stamp forming of both com-modity TPCs [8]and high performance TPCs [9]. Compressive fiber stresses originate from bending of the laminates in this process. Lami-nate bending may also cause waviness in TSC parts that are produced with hot drape forming[10]or double diaphragm forming[11]. Im-portant associated aspects are ply-ply friction and the lay-up sequence, as they have a large influence on the formation of wrinkles for TSCs [12]and TPCs[13]. Nowadays macroscale wrinkling can be predicted reasonably well [14]and software for wrinkle prediction is commer-cially available [15]. However, the simulation results are limited to large wrinkles[13]. It was demonstrated that the prediction of wrinkle formation in forming processes on the scale of a fiber or tow is in fact possible[16]. However, the authors used indirect indicators such as the shear angle, while a reliable prediction requires models based on the underlying physics.

Mesoscale buckling can occur due to compressive fiber stresses in TSC processing when a stack of plies is debulked over a convex surface [17], during lay-up with automated fiber placement (AFP) [18], or during curing of stepped laminates, tapered laminates and laminates with gaps and overlaps from the AFP process, shown by modelling[19] and with experiments [20]. Similar studies for TPCs have not been reported in the literature.

Tool-ply interaction is another mechanism reported in the literature to cause an axial compressive stress in the fiber reinforcement on the mesoscale. A laminate can be strained when the coefficient of thermal expansion (CTE) of the tool is higher than the CTE of the fiber re-inforcement. In TSCs this CTE mismatch may lead to fibers being stretched by the hot tool during curing, which cause shape distortions at room temperature [7]. Conversely, a CTE mismatch may lead to compressive fiber stresses by the shrinking tool in TPC processing, causing waviness. This was shown by observing fiber waviness after the production of T300/P1700 carbon fiber/polysulfone (C/PSU) laminates with a fiber volume fraction (Vf) of 60%, using tools with a varying CTE [21]. The use of tool plates with a high CTE yielded wavy laminates, while tool plates with a low CTE produced no observable waviness. It has been hypothesized that waviness must form while cooling through the glass transition temperature, since the processing temperature had no influence on waviness severity in C/PSU laminates[22]. C/PSU is a TPC with an amorphous matrix. It is unknown if the same conclusions hold for semi-crystalline TPCs, as they have a different solidification mechanism. The influence of tool-ply interaction on the formation of waviness in TPCs with a semi-crystalline matrix has not yet been re-ported in the literature.

In TSC processing, matrix shrinkage is often considered a source of axial compressive fiber stresses on the microscale. Researchers have

modelled a single fiber embedded in thermoset resin during curing and concluded that matrix shrinkage may cause a fiber to buckle[23]. This effect has also been observed in-situ[24]. Buckling of a single fiber is highly dependent on the temperature evolution[25]. However, results from the single fiber experiments research are not directly transferable to laminate production. The stabilization surrounding a fiber in a ply is very different from a single fiber embedded in resin, because of the high Vf and fiber–fiber interaction[26]. Matrix shrinkage is a suggested mechanism for waviness formation in TPCs[27], although no evidence for this mechanism has been reported in the literature.

Axial compressive fiber stresses are more prone to cause waviness when a laminate has deconsolidated or decompacted. This can occur during the heating phase of the stamp forming process[28]. Similarly, TPC prepreg tapes desonsolidate during the heating phase of AFP lay-up [29]. Moisture release from the matrix and the release of residual stresses have been identified as the underlying mechanisms.

1.2.2. Transverse fiber stress

The other loading condition for waviness formation is a stress transverse to the fiber. The prominent mechanism reported in the lit-erature is flow of material. Resin flow, which is inherent to the resin transfer molding (RTM) process for TSCs, may produce waviness in the fiber reinforcement[30]. Flow of material during production of lami-nates may originate from a pressure gradient during procesing[31]. In TPC processing these can be caused by an improper thickness dis-tribution or mold cavity mismatch in the case of matched mold con-solidation[32]. However, transverse fiber stresses are not expected to play a role in the processes considered in this research.

1.3. Objective

Full control of the process is needed to reduce the occurrence of waviness in TPC parts. Therefore, the objective of this research is to describe the governing mechanisms for waviness formation in semi-crystalline TPCs during flat laminate consolidation. This is a pre-requisite for the development of a predictive model. Furthermore, the relevant temperature interval for waviness formation must be de-termined, as it is yet unclear at which temperatures waviness is formed. This enables characterization of the required material properties for modeling purposes and leads to identification of the processing para-meters that could be improved for the production of waviness-free la-minates. This may readily lead to processing guidelines for minimiza-tion of waviness formaminimiza-tion.

2. Experimental work

The proposed waviness formation mechanisms mentioned in lit-erature could only be deducted a posteriori since observation of an actual part during processing has not been performed. Therefore, an Fig. 1. Examples of in-plane fiber waviness. Left: Flat carbon fiber/PEEK composite, press consolidated. Right: Double curved carbon fiber/PEEK part, stamp formed. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

(3)

experimental setup has been designed to visually observe the formation of waviness during consolidation of flat TPC laminates. The physical mechanisms governing the formation of waviness may possibly be de-duced from the visual observation.

2.1. Description of the vacuum assisted glass plate consolidation setup The setup is built on a custom hotplate designed for vacuum bag consolidation of TPCs. An overview of the setup is shown inFig. 2. An aluminum thermal diffuser plate is placed on top of the hotplate, fol-lowed by a graphite sheet and an aluminum caul sheet. The graphite sheet provides a slip plane between the aluminum heat diffuser and the caul sheet. This slip plane ensures that the thermal expansion of the aluminum caul sheet is not constrained by the layers below the caul sheet. Aluminum was selected as the caul sheet material for its high thermal expansion, which is associated with waviness formation[21]. The temperature was measured next to the glass plate using two ther-mocouples. Cooling was provided by compressed air flowing through cooling channels in the hotplate. Pressure was applied to the laminate by a vacuum bag. A small square was cut out of the vacuum film to create an observation window, which enables real-time observation of the composite. A camera was used to record the composite during processing. Details about these parts are listed inTable 1.

The glass plate was treated with Frekote NC700, a releasing agent designed for a low friction surface during processing[33,34]. A Upilex 25S polyimide film was placed between the caul sheet and the com-posite. This film enables release of the composite from the caul sheet after processing, while retaining a high tool-ply friction during pro-cessing. This effect has been validated with the friction measurement device as described in[35]and shown inFig. 3. A coefficient of Cou-lomb friction of 0.42 was found between the caul sheet and the com-posite, for a temperature of 385°C and a pressure of 1 bar. The high friction ensured that the composite followed the thermal expansion of the caul sheet, while moving freely along the glass surface.

2.2. Materials

The caul sheet is a 1.5 mm thick aluminum 1050 alloy sheet with a CTE of 22·10 6K−1at room temperature. The glass plate is an 11 mm thick borosilicate plate with a CTE of 3.3·10 6K−1[36]. The composite material is C/PEEK, a unidirectional (UD) carbon fiber thermoplastic composite prepreg tape with a Vfof 59%. C/PEEK UD prepregs were supplied by Toray Advanced Composites (supplier T)[37]and Solvay (supplier S)[38]. Both prepregs contain AS4 fibers, with a small ne-gative CTE of 0.63·10 6K−1[39]. The PEEK matrix in both prepregs has a glass transition temperature (Tg) of 143 °C and a melting point (Tm) of 343 °C. The CTE of neat PEEK resin is 55·10 6K−1at room temperature and 140·10 6 K−1in melt[40]. Prepregs from different suppliers were tested to investigate the effect of a difference in fi-ber–matrix distribution between the material from these suppliers[29], since a resin rich surface can potentially decrease the tool-ply interac-tion through a lubricainterac-tion mechanism[41]. A cross section of both prepregs is shown inFig. 4, where the differences in fiber–matrix dis-tribution and the resin-rich layer are clearly visible between the two materials.

2.3. Method

Two types of experiments were performed. The first set of experi-ments was designed to observe the consolidation of the laminate using the setup shown inFig. 2. The second series of experiments was de-signed to study the effect of processing conditions on wrinkle forma-tion, using prepregs from both suppliers. Here, the glass plate was re-placed by an aluminum caul sheet.

Fig. 2. Overview of the experimental setup. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Table 1

Details about the parts used in the experimental setup.

Component Details

Hotplate 400 × 400 mm Brass plate, 4 kW heating rods Cooling Cooling channels with compressed air Aluminum diffuser Aluminum 1050 alloy

Vacuum sealant Airtech A800-3G

Vacuum film Airtech Thermalimide E 50 µm Graphite plate Egraflex NR, 1 mm

Caul sheet Aluminum 1050 alloy, 304 × 304 × 1.5 mm Release media UBE Upilex 25S

Composite C/PEEK, 0.15 mm UD ply, 59 %Vf Glass plate Duran borosilicate 11 mm thickness Glass breather Airtech UHT 800

Camera IDS uEye UI549xSE-M

(4)

2.3.1. Real-time observation of the laminate surface during consolidation Three single ply laminates of 300 mm x 300 mm were consolidated with the setup as described in Section 2.1, using the Toray Cetex TC1200 prepreg material without any pre-treatment. The following procedure was used as a baseline and is visualized inFig. 5.

1. Apply vacuum to generate a pressure difference of 1 bar 2. Heat to Tmax(385 °C) at a rate of 8 °C/min

3. Dwell for 10 min at Tmax

4. Cool to room temperature at a rate of 8 °C/min 5. Release vacuum and demould

The temperature was logged throughout the cycle and a picture from the camera was saved every 10 s. A time-lapse video was com-posed from the pictures with their corresponding temperature values embedded. Therefore, detection of waviness formation and other visual events becomes straightforward and the waviness formation tempera-ture interval could be determined.

2.3.2. Influence of processing temperature and transverse plies

Kugler et al. observed that the maximum process temperature did not influence waviness formation for amorphous TPCs[22]. Therefore, experiments were performed to investigate if this also holds for semi-crystalline TPCs. The glass plate has been replaced with an aluminum caul sheet as visual observation is not necessary in these additional experiments. This ensured an equal tool-ply interaction on both sides of the laminate. The maximum temperature (Tmax) was varied between 360–400 °C. Single ply [0] and four-ply [0/90]slaminates were pro-cessed to investigate the effect of transverse plies in the lay-up. Prepregs from two different suppliers were tested to rule out supplier specific effects. No pre-treatments were performed on the material. The test matrix is available inTable 2. Three laminates were processed for each combination of parameters.

Surface micrographs were taken for nine patches of wavy fibers, evenly selected from each laminate, using a Keyence VHX-5000 mi-croscope at a magnification of 200x. The micrographs were processed with a fiber tracing algorithm to determine the wavelength and max-imum misalignment angle for the wavy fibers[42]. These wave para-meters have been correlated to compressive strength[5], hence they quantify the severity of the produced waviness. Cross sectional micro-graphs of wavy patches were taken using aforementioned microscrope, to study the through thickness propagation of the waviness for the four-ply laminates.

3. Results

3.1. Real-time observation of the laminate surface during consolidation Four pictures of key moments during the consolidation of a single ply are shown inFig. 6with their according temperature. The following observations were made:

(Fig. 6.1 6.2) No waviness formation was observed during heating from room temperature to Tmax.

(Fig. 6.2 6.3) Formation of waviness was observed during cooling down from Tmax. New waves were formed and existing waves gradually increased in severity. The movement of fibers had completely stopped at approximately 285 °C. The reflection of light on the ply changed around this temperature.

(Fig. 6.3 6.4) During cooling from 285 °C down to room tem-perature, no more formation of waviness was observed while existing waves did not change in shape anymore.

Waviness appeared at all places in the observation window, despite it being only visible at the edges in these pictures. This is due to diffi-culties with capturing waviness with the camera, since observation is only possible from a specific angle related to the incoming light. 3.2. Influence of processing temperature and transverse plies

The resulting wave parameters for all laminates that were con-solidated between two caul sheets are shown inTable 2. Each experi-ment was performed three times and nine patches of wavy fibers were analyzed per laminate, to obtain the wavelength ( ) and maximum misalignment angle ( ). The averaged wave parameters (¯, ¯) were determined by averaging all wavy patches in an experiment.Fig. 7 Fig. 4. Typical cross sections of the prepreg material before processing. Top: Solvay AS4/APC2 with a rough surface and resin-rich layer on the top side.

Bottom: Toray Cetex AS4/TC1200 with a smooth

surface.

Fig. 5. Temperature profile during consolidation.

Table 2

Test matrix and wave parameters for wavy patches found in the bottom ply of laminates that were consolidated between two aluminum caul sheets. # = Experiment nr, S&L = supplier and lay-up, = wavelength, = maximum misalignment angle and ± indicates the standard deviation. ¯ and ¯ are aver-aged values for all wavy patches in the corresponding experiments.

# S & L Tmax ¯ ¯ [°C] [mm] [°] [mm] [°] 1 T [0] 360 3.40±0.26 9.66±1.37 3.30±0.41 9.75±1.48 ± 3.45 0.52 9.06±1.53 ± 2.99 0.30 10.69±1.00 2 T [0] 385 4.19±0.62 15.44±2.69 4.12±0.53 15.14±2.55 ± 4.19 0.55 15.67±2.55 ± 3.95 0.35 14.15±2.09 3 S [0] 385 4.91±0.68 15.42±1.03 4.91±0.62 15.30±1.89 ± 4.73 0.69 15.77±1.86 ± 5.09 0.37 14.69±2.37 4 T [0] 400 4.34±0.35 17.84±2.52 4.02±0.45 17.18±2.01 ± 3.78 0.41 16.83±1.48 ± 3.84 0.31 16.74±1.33 5 T [0/90]s 385 3.58±0.28 13.60±1.75 3.82±0.46 14.02±1.59 ± 3.98 0.53 14.47±1.77 ± 3.92 0.41 13.88±0.63

(5)

shows the surface of a four-ply laminate in which multiple patches of wavy fibers are visible. Although the picture may suggest otherwise, all patches solely exhibit in-plane waviness. The surface micrograph (outlined in red) is processed with the fiber tracing algorithm. All la-minates exhibit fiber waviness in a similar way: patches of wavy fibers, scattered all over the surface with a waviness-free margin of approxi-mately 3 cm from the edge, as illustrated in Fig. 8. These waves re-semble sinusoids consisting of 1,5–2 periods. A bundle of fibers typi-cally passes through three to five wavy patches when it is traced from one edge of the laminate to the other edge. The difference between the curvilinear length and span length of a fiber in one patch of waviness is approximately 0.15 mm.

Cross-sectional micrographs of a wavy patch in a four-ply laminate are shown inFig. 9. The top image shows a cross section parallel to the

surface fiber direction, where waviness is clearly visible in the bottom layer. Although waviness appeared in the outer ply on both sides of these laminates, the location of a wavy patch on one side of the lami-nate does not necessarily match the location of a wavy patch on the other side of the laminate. The bottom image shows a cross section perpendicular to the surface fiber direction. No waviness was observed in the middle plies.

The results for the experiments can be compared inFig. 10, based on the results inTable 2. No significant differences in misalignment angle have been observed between the wavy patches in a single-ply and a four-ply laminate, and between both manufacturers. A significant in-crease in misalignment angle was observed with an increasing Tmax between 360–400 °C. No clear trend was observed for the wavelength. Fig. 6. Pictures of the observation window during a consolidation cycle of a single ply C/PEEK. (1) Start of cycle. (2) Maximum temperature, start of waviness formation during cooling. (3) End of waviness forma-tion during cooling. (4) End of cycle. Please note that the edges of the observation window as illustrated inFig. 2are shown, not the laminate edges.

Fig. 7. Background: Picture of the surface of a four-ply laminate (#5 inTable 2). Patches of wavy fibers are clearly visible.

Fore-ground: Surface micrograph with typical

waviness as found by the experiments. A traced fiber is shown as a bold line. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

(6)

4. Discussion

4.1. Real-time observation of consolidation through glass plate

Waviness formation was only observed during the cooling stage between Tmaxand 285 °C. Waviness formation stopped at around 285 °C in all experiments. DSC measurements were performed using a Mettler Toledo DSC822E differential scanning calorimeter to explain this. Three samples of approximately 15 mg C/PEEK prepreg were analyzed in a nitrogen environment, to prevent degradation. The temperature profile fromFig. 5was used.

The traces of the DSC measurements are shown in Fig. 11. This graph shows that crystallization of the C/PEEK composite took place between 305–285 °C. The latter temperature is the crystallization endset (Tcryst), which corresponds with the change in reflection of light on the ply during the observation experiments. This is a good estimate for the temperature where the matrix has fully solidified[43].

The difference between Tmaxand Tcrystwill be called the waviness formation temperature interval:

=

Tw Tmax Tcryst (1)

4.2. Waviness formation mechanisms in the waviness formation temperature interval

An overview of the possible waviness formation mechanisms was given in Section1.2. Deconsolidation and material flow have not been observed in the experiments. The remaining mechanisms are discussed next.

4.2.1. Matrix shrinkage

The matrix is in a molten, viscous state between Tmaxand Tcryst. The storage modulus and viscosity of the matrix are low in this interval. Compressive fiber stresses arise from the CTE mismatch between the matrix and the fiber during cooling. However, when the matrix is in a viscous state, it is able to relax its internal stresses to some extent. Consequently, compressive fiber stresses are relaxed as well. Stress re-laxation is much more difficult when the matrix is crystalline. Therefore, it is argued that if compressive fiber stresses are caused by differential shrinkage of the fiber and the matrix, they would primarily arise when the matrix is crystalline. However, no waviness formation has been observed in the experiments after crystallization.

Furthermore, no increased waviness formation or intensification has been observed during crystallization, when significant shrinkage of the matrix is taking place[44]. Hence, the contribution of matrix shrinkage to the formation of fiber waviness is considered negligible.

4.2.2. Tool shrinkage

The aluminum caul sheets expand during heating and shrink during cooling. With the presence of tool-ply friction and a tool-ply CTE mis-match, the fibers will experience a tensile force during tool expansion. However, this tensile fiber force is limited by tool-ply slip when the maximum attainable tool-ply friction force is reached. During cooling the tool shrinks, imposing a compressive stress on the fibers. These compressive fiber stresses are the presumed driving force for fiber buckling and the observed patches of waviness. Evidence is found in the experimental results shown inFig. 10. A higher Tmaxleads to more tool shrinkage, which leads to a larger misaligment angle. The cross sections inFig. 9show that the waviness is a surface effect: it does not propagate through the thickness in a [0/90]slaminate. Furthermore, the amount of shrinkage for the 300 mm wide caul sheet is 0.66 mm when con-sidering its linear CTE of 22·106K−1and a T

wof 100 °C. The dif-ference between the curvilinear length and span length of a fiber in a patch of waviness was approximately 0.15 mm. Since a fiber typically passes through three to five wavy patches, the total length difference is between 0.45–0.75 mm, which is roughly equal to the tool shrinkage.

Therefore, the conclusion in the literature, that Tmaxhas no influ-ence on waviness formation in amorphous TPCs[22], is not applicable to semi-crystalline TPCs. The absence of crystallization in amorphous TPCs leads to a different solidification mechanism and a much more gradual increase of matrix viscosity and storage modulus during cooling.

4.2.3. Buckling resistance of the fiber–matrix system

The storage modulus of the matrix dictates the transverse stabili-zation of the fibers. A higher modulus provides a better quasi-static buckling resistance[45]. During Twthe storage modulus of the matrix is low, hence the buckling resistance is low. At Tcrystthe matrix modulus is high, therefore the buckling resistance is also high.

Two situations occur during cooling: Below Tcryst, the buckling re-sistance is high and compressive fiber forces due to matrix shrinkage are high. Above Tcrystthe matrix is in melt: the buckling resistance is low and compressive fiber stresses can be dissipated through the vis-cous matrix. In the experiments described in this paper, waviness Fig. 8. Illustration of typically observed waviness in the experiments: patches of wavy fibers scattered over the surface of a laminate.

Fig. 9. Cross-sectional micrographs of a wavy patch located on the bottom ply, found in a [0/ 90]s laminate that was consolidated between two aluminum caul sheets. Top: Cross-section parallel to the surface fiber direction. Waviness is clearly visible in the bottom ply. Bottom: Cross-section perpendicular to the surface fiber direc-tion.

(7)

formation is only observed above Tcrystwhere the buckling resistance of the fiber–matrix system is low. Therefore it is argued that formation of waviness in semi-crystalline TPCs takes place during cooling, before crystallization, with tool shrinkage imposing a compressive axial force on the fibers. This statement motivates further investigation into the role of tool-ply friction and tool material CTE on the formation of fiber waviness during semi-crystalline TPC processing. Improving these as-pects of the process could minimize or eliminate fiber waviness in TPC laminate consolidation. Ongoing experiments using low CTE Kovar tool plates yielded no waviness, which is in line with the proposed me-chanisms. A more elaborate investigation concerning multiple tool materials, tool-ply friction coefficients and laminate lengths will be submitted for publication in due time.

5. Conclusions

An experimental setup for real-time observation of waviness for-mation during the production of thermoplastic composite laminates has been presented in this paper. It was successfully used to produce la-minates with patches of wavy fibers, to observe the formation of wa-viness and to determine the temperature interval in which wawa-viness formation takes place in C/PEEK prepregs.

It can be concluded that fiber waviness in TPCs forms during cooling between the maximum process temperature (Tmax) and the crystal-lization temperature (Tcryst). This temperature interval is called the waviness formation temperature interval and has been defined as ΔTw= Tmax– Tcryst. It should be mentioned that the waviness severity did not increase below Tcrystand no new waviness formation was ob-served. One key observation was that the maximum misalignment angle of the wavy fibers increased when Twincreases.

The driving force for waviness formation during C/PEEK con-solidation was determined to be an axial compressive stress on the fiber, caused by tool displacements that are transferred to the laminate via tool-ply friction. The waviness can be considered a surface

phenomenon, as it was not observed in plies below the surface. Matrix dominated mechanisms, such as matrix shrinkage, were considered negligible since the formation of fiber waviness was only observed in the temperature interval with a low matrix storage modulus.

The determined waviness formation temperature interval and gov-erning mechanisms can now be used in a modelling approach, with the objective of producing guidelines for waviness-free laminates. The tool-ply friction, the coefficient of thermal expansion of the tool material and the maximum processing temperature were the identified para-meters that can be altered to decrease the waviness severity.

CRediT authorship contribution statement

E.T.M. Krämer: Conceptualization, Methodology, Investigation,

Writing - original draft. W.J.B. Grouve: Conceptualization, Supervision, Writing - review & editing. S. Koussios: Resources, Writing - review & editing. L.L. Warnet: Conceptualization, Supervision. R. Akkerman: Project administration, Writing - review & editing.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influ-ence the work reported in this paper.

Acknowledgements

The authors gratefully acknowledge the financial and technical support from the industrial and academic members of the ThermoPlastic composites Research Center (TPRC), as well as the support funding from the Province of Overijssel for improving the re-gional knowledge position within the Technology Base Twente in-itiative.

Fig. 10. Average wavelength and maximum misalignment angle. Each bar represents the corresponding averaged wavelength ( ¯) and angle (¯) fromTable 2. T = Toray, S = Solvay. Colors correspond with Tmax, error bars indicate the standard deviation. The temperature in the legend is Tmax. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 11. Relevant part of the DSC heating and cooling trace for three C/PEEK samples. The temperature profile fromFig. 5is used in these experiments. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

(8)

References

[1] Rosen B. Mechanics of composite strengthening. Fibre Compos Mater 1965:37–75. [2] Budiansky B, Fleck NA. Compressive failure of fibre composites. J Mech Phys Solids

1993;41(1):183–211.https://doi.org/10.1016/0022-5096(93)90068-Q. [3] Bloom LD, Wang J, Potter KD. Damage progression and defect sensitivity: an

ex-perimental study of representative wrinkles in tension. Compos Part B: Eng 2013;45(1):449–58.https://doi.org/10.1016/j.compositesb.2012.05.021. [4] Wisnom MR, Atkinson J. Fibre waviness generation and measurement and its effect

on compressive strength; 2000. doi:10.1177/073168440001900201.

[5] Piggott MR. The effect of fibre waviness on the mechanical properties of unidirec-tional fibre composites: A review. Compos Sci Technol 1995;53(2):201–5.https:// doi.org/10.1016/0266-3538(95)00019-4.

[6] Highsmith AL, Davis JJ, Helms KL. Influence of fiber waviness on the compressive behavior of unidirectional continuous fiber composites. ASTM Spec Tech Publ 1992(1120):20–36.https://doi.org/10.1520/stp20142s.

[7] Potter KD. Understanding the origins of defects and variability in composites manufacture. 17th International conference on composite materials. 2009. p. 27–31.

[8] Vanclooster K, Lomov SV, Verpoest I. On the formability of multi-layered fabric composites. ICCM International conferences on composite materials. 2009. [9] McGuinness GB, Brádaigh CMÓ. Effect of preform shape on buckling of

quasi-iso-tropic thermoplastic composite laminates during sheet forming. Compos Manuf 1995;6(3–4):269–80.https://doi.org/10.1016/0956-7143(95)95020-Y. [10] Hallander P, Akermo M, Mattei C, Petersson M, Nyman T. An experimental study of

mechanisms behind wrinkle development during forming of composite laminates. Compos Part A: Appl Sci Manuf 2013;50:54–64.https://doi.org/10.1016/j. compositesa.2013.03.013.

[11] Gutowski TG, Dillon G, Chey S, Li H. Laminate wrinkling scaling laws for ideal composites. Compos Manuf 1995;6(3–4):123–34. https://doi.org/10.1016/0956-7143(95)95003-H.

[12] Hallander P, Sjölander J, Åkermo M. Forming induced wrinkling of composite la-minates with mixed ply material properties; an experimental study. Compos Part A: Appl Sci Manuf 2015;78:234–45.https://doi.org/10.1016/j.compositesa.2015.08. 025.

[13] Haanappel S. Forming of UD fibre reinforced thermoplastics Ph.D. thesis University of Twente; 2013.

[14] Haanappel SP, ten Thije R, Akkerman R. Forming predictions of UD reinforced thermoplastic laminates. European Conference on Composite Materials (ECCM14). 2010.

[15] Aniform Engineering B.V., Aniform – Virtual Forming Software; 2019.www. aniform.com.

[16] Harrison P, Gomes R, Curado-correia N. Composites: Part A Press forming a 0/ 90 cross-ply advanced thermoplastic composite using the double-dome benchmark geometry. Compos Part A 2013;54:56–69.https://doi.org/10.1016/j.compositesa. 2013.06.014.

[17] Dodwell T, Butler R, Hunt G. Out-of-plane ply wrinkling defects during consolida-tion over an external radius. Compos Sci Technol 2014;105:151–9.https://doi.org/ 10.1016/j.compscitech.2014.10.007.

[18] Beakou A, Cano M, Cam JL, Verney V. Modelling slit tape buckling during auto-mated prepreg manufacturing: a local approach. Compos Struct

2011;93(10):2628–35.https://doi.org/10.1016/j.compstruct.2011.04.030. [19] Li X, Hallett S, Wisnom M. Modelling the effect of Gaps and Overlaps in Automated

Fibre Placement (AFP) manufactured laminates. Sci Eng Compos Mater 2015;22(2):115–29.https://doi.org/10.1515/secm-2013-0322.

[20] Hallett SR, Belnoue JP-H, Nixon-Pearson OJ, Mesogitis T, Kratz J, Ivanov DS, et al. Understanding and Prediction of Fibre Waviness Defect Generation. American so-ciety for composites: thirty-first technical conference. 2016.

[21] Bhalerao MS. On process-induced fiber waviness in composites: theory and ex-periments, Ph.D. thesis; 1996.

[22] Kugler D, Moon TJ, Engineering M, Austin A. Identification of the most significant

processing parameters on the development of fiber waviness in thin laminates. J Compos Mater 2002;36(12):1451–79.https://doi.org/10.1106/

002199802023575.

[23] Jochum C, Grandidier JC. Microbuckling elastic modelling approach of a single carbon fibre embedded in an epoxy matrix. Compos Sci Technol

2004;64(16):2441–9.https://doi.org/10.1016/j.compscitech.2003.12.004. [24] Jochum C, Grandidier JC, Smaali MA. Experimental study of long T300 carbon fibre

undulations during the curing of LY556 epoxy resin. Compos Sci Technol 2007;67(11–12):2633–42.https://doi.org/10.1016/j.compscitech.2006.01.006. [25] Jochum C, Grandidier JC, Smaali M. Proposal for a long-fibre microbuckling

sce-nario during the cure of a thermosetting matrix. Compos Part A: Appl Sci Manuf 2008;39(1):19–28.https://doi.org/10.1016/j.compositesa.2007.09.012. [26] Lightfoot JS. Mechanisms of defect formation in carbon fibre composites Ph.D.

thesis University of Bristol; 2013.

[27] Parlevliet PP, Bersee HEN, Beukers A. Residual stresses in thermoplastic composites – a study of the literature. Part III: E ff ects of thermal residual stresses. Compos Part A: Appl Sci Manuf 2007;38:1581–96.https://doi.org/10.1016/j.compositesa.2006. 12.005.

[28] Slange TK, Warnet LL, Grouve WJ, Akkerman R. Deconsolidation of C/PEEK blanks: on the role of prepreg, blank manufacturing method and conditioning. Compos Part A: Appl Sci Manuf 2018;113(June):189–99.https://doi.org/10.1016/j. compositesa.2018.06.034.

[29] Kok T. On the consolidation quality in laser assisted fiber placement: the role of the heating phase. Ph.D. thesis University of Twente; 2018.https://doi.org/10.3990/1. 9789036546065.

[30] Brocks T, Shiino MY, Cioffi MOH, Voorwald HJC, Caporalli Filho A. Experimental RTM manufacturing analysis of carbon/epoxy composites for aerospace application. Mater Res 2013;16(5):1175–82.

https://doi.org/10.1590/S1516-14392013005000107.

[31] Chan WS, Wang JS. Influence of fiber waviness on the structural response of composite laminates. J Thermoplast Compos Mater 1994;7(3):243–60.https://doi. org/10.1177/089270579400700306.

[32] Slange TK. Rapid manufacturing of tailored thermoplastic composites by automated lay-up and stamp forming a study on the consolidation mechanisms; 2019. [33] Henkel, Loctite Frekote 700NC – Technical Data Sheet; 2015.

[34] Murtagh AM, Lennon JJ, Mallon PJ. Surface friction effects related to pressforming of continuous fibre thermoplastic composites. Compos Manuf 1995;6(3–4):169–75.

https://doi.org/10.1016/0956-7143(95)95008-M.

[35] Sachs U. Friction and bending in thermoplastic composites forming processes. Ph.D. thesis University of Twente; 2014.https://doi.org/10.3990/1.9789462594838. [36] Schott, Borosilicate glass 3.3 - Technical Data Sheet; 2016.

[37] Toray Advanced Composites, Cetex TC1200 PEEK – Technical Data Sheet; 2019. [38] Solvay, APC-2 PEEK Thermoplastic Polymer Prepreg – Technical Data Sheet; 2017. [39] Hexcel, HexTow AS4 Carbon Fiber – Technical Data Sheet; 2018.

[40] Victrex High Performance Polymers, PEEK 150G/151G datasheet; 2017. [41] ten Thije RH, Akkerman R, van der Meer L, Ubbink MP. Tool-ply friction in

ther-moplastic composite forming. Int J Mater Form 2008;1(SUPPL. 1):953–6.https:// doi.org/10.1007/s12289-008-0215-9.

[42] Krämer E, Koussios S, Warnet LL, Grouve W, Akkerman R. Automatic fiber tracing and waviness detection in carbon fiber composite micrographs. In: SAMPE Europe conference 2019, Nantes; 2019.

[43] Brauner C, Peters C, Brandwein F, Herrmann AS. Analysis of process-induced de-formations in thermoplastic composite materials. J Compos Mater

2014;48(22):2779–91.https://doi.org/10.1177/0021998313502101.

[44] Nairn JA, Zoller P. The development of residual thermal stresses in amorphous and semicrystalline thermoplastic matrix composites. Toughened composites West Conshohocken, PA: ASTM International; 1987. p. 328–41.https://doi.org/10. 1520/STP24386S.

[45] Kassapoglou C. Beam on an elastic foundation under compression. Design and analysis of composite structures. 2013. p. 200–5.

Referenties

GERELATEERDE DOCUMENTEN

Hooggerechtshof was het gebrekkige contact reden om aan te nemen dat zich tussen de vader en zoon Kürsad geen gezinsleven had ontwikkeld, terwijl de positieve verplichting

In addressing the question: “How can the concept of affect provide us with a theoretical framework from which we can approach bioart?”, I have demonstrated that a Deleuzian reading

Optreden tegen de piraten in dit gebied was lastig om twee redenen: ten eerste werd het Indonesische militaire apparaat niet volledig betaald door de overheid, maar

For three-dimensional isotro- pic assemblies the macro-scale elastic characteristics are described by the bulk and the shear modulus, which depend on the micro-scale characteristics

Het boek van Davis en Hersh behandelt een enorme lijst van onderwerpen die de wiskundige herkent als behorend bij zijn 'cultuur': 'waarom werkt wiskunde?', 'het nut van de

The objective of this study was to compare the AmpliChip CYP450 Test W (AmpliChip) to alternative genotyping platforms for phenotype prediction of CYP2C19 and CYP2D6 in a

Ainsi, ce retranchement semble avoir été doté sur toute sa longueur d'un front palissadé qui a été remplacé à deux reprises et a été muni de deux larges entrées

dichloride, adsorbed on pyrolytic graphite, reduction to water occurred in alkaline solution, whereas in acid solution only hydrogen peroxide was formed; the results are compared