Lamellar structure obtained by bringing the surface of a molten plaque into contact with a metal plate held at room temperature (microstructure just behind the surface) 

Lamellar structure obtained by bringing the surface of a molten plaque into contact with a metal plate held at room temperature (microstructure just behind the surface) 

Source publication
Article
Full-text available
The crystallization kinetics of polyamide-12 has been investigated using a combination of differential scanning calorimetry (DSC) and hot-stage optical microscopy. The DSC data for isothermal crystallization were consistent with a simple two-parameter Avrami model for isothermal crystallization and optical measurements of the spherulite growth rate...

Context in source publication

Context 1
... ®t of Eq. (11) to the data in Fig. 4 gave s o 1.5 ́ 10 ) 4 s and K N 110,000 K 2 (cf. Eq. 10). The form of Eq. (11) was chosen somewhat arbitrarily, but it was found to account well for the observed temperature dependence of s . (cf. the classical Turnbull±Fisher expression for primary nucleation, for example, in which the argument of the second exponential is raised to the power 2 [20]). The data obtained from nonisothermal crystallization at 40 Kmin ) 1 are compared in Fig. 9 with curves calculated using parameters derived from the isothermal data (Fig. 6, Eqs. 9, 11) and a simulation made after adjusting N o and T sd to optimize the ®t. Also shown is the optimized ®t plotted as a function of the sample temperature, T s , and the corresponding degree of conversion as a function of T s . This demonstrates that crystallization itself took place over a relatively narrow range of sample temperatures between 140 and 146 ° C, in spite of the fact that the oven temperature, T f , corresponding to the end of the crystallization exotherm was about 132 ° C. For nominal cooling rates above 40 Kmin ) 1 adequate control of the temperature ramp became dicult [9] and 146 ° C was therefore considered to be the lower limit of the temperature range which could be investigated in detail. DSC results for various cooling rates are shown in Fig. 10 along with the calculated curves. The adjusted values of N o ( T sd ) used to obtain these curves are compared with the results from isothermal experiments in Fig. 11 and the adjusted values of T sd are compared with the predictions of Eqs. (9) and (11) in Fig. 12. The dierences between the predicted and adjusted values of T sd were comparatively small. The dierences between the predicted N o ( T sd ) and adjusted values of N 0 were more signi®cant, as borne out by Fig. 9. However the predicted values relied on a relatively long extrapolation of data points con®ned to a narrow range of temperatures in which N o was changing relatively slowly with T . An exponential temperature dependence has been observed in other polymers [10, 11], but given the assumptions implicit in the use of Eq. (10) for G , corroborating evidence for the evolution in N o from optical microscopy was considered to be important. The results of direct estimates of the average spherulite radii are shown in Fig. 13, where they are compared with the radius of a sphere with volume N o À 1 . The agreement is satisfactory, although in the case of the sample cooled at 40 Kmin ) 1 it was dicult to characterize the morphol- ogy precisely, owing to overlap in the projections of individual spherulites and an ill-de®ned spherulitic texture. Time±temperature transformation diagrams are a con- venient way of representing either data or models for solidi®cation. The isothermal time±temperature trans- formation diagram for PA-12 inferred from the present model and extrapolated to higher undercooling using N o 9 ́ 10 41 exp( ) 0.153 T ) ( T in Kelvin) to interpolate and extrapolate N o ( T ) (the hatched curve in Fig. 11) is shown in Fig. 14. It should be emphasized that the induction time s is not expected to correspond to a true onset since it is derived from a ®t of a two-parameter Avrami model to data from a system in which the eective induction times may be distributed (and in which thermal nucleation may also be present). On the other hand, the curves for 50 and 99% conversion are realistic for the temperature range in which data were available. Figure 15a shows the non-isothermal time±temperature conversion diagram for a starting temperature of 220 ° C, chosen to be consistent with the DSC data (a dierent choice of starting temperature would simply result in a shift in the zero of the time axis). The curves were calculated from Eq. (6) and therefore do not include eects of latent heat evolution or thermal lag and will therefore, in general, not re ̄ect the apparent behaviour in practical situations. Indeed, as indicated in Fig. 15a, the predicted time for 99% conversion diverged from the experimental DSC data quite mark- edly at the highest cooling rates, even when these latter were corrected for thermal lag. Nevertheless certain situations, such as quenching against a cold metallic mould, will be approximately consistent with the ideal case. Also shown for completeness in Fig. 15b is a time±temperature degree of crystallinity diagram, tak- ing the enthalpy of crystallization of 100% crystalline sample to be 130 Jmol ) 1 [19] and using the data shown in Fig. 2. The question naturally arises as to the extent to which the extrapolation of the present model to lower crystallization temperatures and/or faster solidi®cation rates in Figs.14 and 15 is justi®ed. The extrapolation of G is arguably physically based [13, 21], and the D T dependence is widely observed, even if reservations have been expressed regarding the form of the transport term. N 0 , on the other hand, was interpolated/extrapolated as- suming a simple exponential dependence on T inferred empirically from the results. There is therefore no a priori basis for assumptions about N o in temperature regimes not accessible here. Also, even if it were possible to justify extrapolation of N o measured at high and intermediate crystallization temperatures, in practice these latter are often strongly dependent on the thermomechanical history (other parameters may also vary, particularly in the presence of a ̄ow ®eld). Therefore, although the application of the present approach to well-controlled DSC experiments has been demonstrat- ed, characterization of the microstructure after crystallization would still be necessary to verify any initial hypothesis regarding the nucleation density in other speci®c cases. In connection with this, the lamellar structures of various samples were investigated brie ̄y. The microstructures of samples crystallized isothermally at dierent temperatures and that of a sample taken from close to the surface of a molten plaque brought into contact with a metallic surface held at room temperature are shown in Figs. 16 and 17. In this latter case, the spherulite envelopes were ill de®ned, although lower-magni®cation images of heavily irradiated samples suggested a mean radius of about 0.5 l m. For comparison, extrapolation of N o 9 ́ 10 exp( ) 0.153 T ) predicted a radius of 5.4 l m at 130 ° C and 3.3 l m at 120 ° C. Estimates of the lamellar thicknesses from the samples shown in Fig. 16 gave values of 7.2, 6.7 and 6.3 nm for crystallization temperatures of 172, 166 and 158 ° C, respectively. If the lamellar thickness is assumed to scale as 1/ D T , the thicknesses expected at 130 and 120 ° C would be 5.7 and 5.65 nm, which is consistent with Fig. 17. Faster cooling by quenching a thin ®lm in ice±water from the melt, on the other hand, led to no discernible features in the TEM or in the optical microscope. There remains the question of the assumption of an induction time given by Eq. (11), which was necessary to account for both the isothermal and the nonisothermal data. However, if the dependence of N o on T is assumed to be a consequence of athermal nucleation, as suggested elsewhere, it is dicult to account for the existence of an induction time at high undercooling [10, 11] and it is at present not possible to justify the form of Eq. (11). It is therefore of interest to consider the consequences of zero induction time. The corresponding crystallization onset as de®ned by 1% conversion is given by the hatched line in Fig. 15a. In this case, the apparent cooling rate necessary for quenching to a predominantly amorphous state is shifted from about 800 Kmin ) 1 to about 1000 Kmin ) 1 . The crystallization kinetics of PA-12 has been success- fully modelled using a simple two-parameter Avrami model for isothermal crystallization, which has also been shown to be consistent with optical measurements of the spherulite growth rates and nucleation density. With small adjustments, parameters inferred from the isothermal DSC data also accounted for DSC data for nonisothermal crystallization, when the dynamic heat balance between the sample and the oven was taken into account. Further work is envisaged with the aim of investigating the balance between heterogeneous and homogeneous nucleation and the signi®cance of the observed incubation times, and the role of the prior thermomechanical history on N ...

Similar publications

Article
Full-text available
Potassium dihydrogen phosphate (KDP) single crystals are the only nonlinear crystals currently used for electro-optic switches and frequency converters in inertial confinement fusion research, due to their large dimension and exclusive physical properties. Based on the traditional solution-growth process, large bulk KDP crystals, usually with sizes...

Citations

... A suspected increase in leak-resistance for laser beam polished sealing surfaces is examined by means of a simple positive pressure leakage test as specified in DIN EN 1779 C [29]. For this purpose, the contact surface between two halves of a hollow specimen Crystallization enthalpy ΔH c [J kg −1 ] 50 [25] is sealed with a rubber gasket and screw-retained with a defined torque. The cavity enclosed is then pressurized via a valve. ...
Article
Full-text available
Purpose Multi Jet Fusion™ (MJF) is a powder bed fusion technology that fuses material locally with infrared radiation. Fabricated parts show high mechanical strength, low porosity and good dimensional accuracy, even for challenging geometries. They do, however, have a distinctive grainy surface with a roughness of Ra > 5 μm limiting possible applications. A fairly recent technology for the smoothing and polishing of polymer surfaces is the surface re-melting by laser beam polishing (LBP). The objective of this study is to characterize the LBP-process of MJF-manufactured Nylon PA12 parts and evaluate it in relation to state-of-the-art finishing methods. Methods PA12 samples produced using the Multi Jet Fusion™ process are laser beam polished with a 30 W CO2-laser system. A numerical model is applied in order to estimate the effects of different processing parameters and laser scanning strategies. Calculated temperature progressions are validated experimentally with thermographic measurements. The laser beam polished surfaces are subsequently characterised by surface topography, tensile strength, surface energy and hardness. Results A re-melting and concurrent smoothing of PA12 surfaces was demonstrated for a number of different processing strategies. The use of a line-focus resulted in the best combination of processing speed, flexibility and surface quality with a reduction in roughness of up to 91%. Conclusion Laser beam polishing is applicable to parts manufactured by Multi Jet Fusion™. Depending on the case of application, it can be preferable to conventional post-processing strategies for fusing residual powder and improving tribological properties.
... With regard to formed spherulitic structures, Plummer et al. (2001) described the positive relationship between the underlying isothermal crystallization temperature and the corresponding spherulite radius and the lamellar thickness [34], consistent with the intralayer dependency of formed microspherulitic structures in non-isothermal laser melting of polyamide 12 [4,6]. The formation of nano-and microspherulitic structures, described in the recent literature, suggests implications for corresponding mechanical properties due to the negative correlation of observed spherulite diameters and the elongation at break, described for the application of isotactic polypropylene [35]. ...
... With regard to formed spherulitic structures, Plummer et al. (2001) described the positive relationship between the underlying isothermal crystallization temperature and the corresponding spherulite radius and the lamellar thickness [34], consistent with the intralayer dependency of formed microspherulitic structures in non-isothermal laser melting of polyamide 12 [4,6]. The formation of nano-and microspherulitic structures, described in the recent literature, suggests implications for corresponding mechanical properties due to the negative correlation of observed spherulite diameters and the elongation at break, described for the application of isotactic polypropylene [35]. ...
Article
Full-text available
Non-isothermal laser-based powder bed fusion (LPBF) of polymers suggests the potential for significantly extending the range of materials applicable for powder-based additive manufacturing of polymers, relying on the absence of a material-specific processing window. To allow for the support-free manufacturing of polymers at a build chamber temperature of 25 °C, applied processing strategies comprise the combination of fractal exposure strategies and locally quasi-simultaneous exposure of distinct segments of a particular cross section for minimizing crystallization-induced part deflection. Based on the parameter-dependent control of emerging cooling rates, formed part morphologies and resulting mechanical properties can be modified. Thermographic in situ measurements allow for correlating thermal processing conditions and crystallization kinetics with component-specific mechanical, morphological, and microstructural properties, assessed ex situ. Part morphologies formed at crystallization temperatures below 70 °C, induced by reduced laser exposure times, are characterized by a nano-spherulitic structure, exhibiting an enhanced elongation at break. An ambient temperature of 25 °C is associated with the predominant formation of a combined (α + γ)-phase, induced by the rapid cooling and subsequent laser-induced tempering of distinct layers, yielding a periodic microstructural evolution. The presented results demonstrate a novel approach for obtaining nano-spherulitic morphologies, enabling the exposure-based targeted adaption of morphological properties. Furthermore, the thermographic inline assessment of crystallization kinetics allows for the enhanced understanding of process-morphology interdependencies in laser-based manufacturing processes of semi-crystalline polymers.
... Studies of the crystallization kinetics of PA 12 are generally done by differential scanning calorimetry (DSC). 6,7 However, the limitation to low cooling rates does not allow DSC to access the crystallization of PA 12 at high supercooling of the melt. ...
Article
Full-text available
Nucleation and crystallization of polyamide 12 (PA 12) have been systematically investigated by fast scanning calorimetry at non-isothermal and isothermal conditions. The critical cooling rates of crystallization and crystal nucleation were determined as 300 and 10,000 K/s, respectively. Moreover, the half-times of nucleation (t1/2,nucl) and overall crystallization (t1/2,cry) show monomodal and bimodal dependencies on the crystallization temperature. t1/2,nucl has an approximate minimum value of about 0.0005 s at 333 K, which is about 10–20 K above the glass transition temperature, and t1/2,cry has two minima of about 0.05 and 0.8 s at about 333 and 383 K, respectively. Comparing the crystallization behavior of PA 12 with other polyamides, the activation energy for crystallization increases and the energy barrier of short-range diffusion decreases with the increase of the amide-group density in the chains. Abstract
... Therefore, studying the isothermal and non-isothermal crystallization kinetics of PA12 is a crucial step in order to understand and control this process. As far we are aware, not many authors have investigated crystallization kinetics of PA12 and all of them focused on the high temperature crystallization which represent the process window in SLS process [2] [3]. The goal of this work is to describe and model the crystallization kinetics of PA12 in the whole temperature range between the glass transition and the melting, to validate the model results with experimental data acquired from combining fast scanning (chip) calorimetry and wide angle X-ray diffraction (WAXD). ...
... A monoclinc structure has been assigned to both and phase although the two WAXD scattering reflections of phase are closer compared to those of phase. Even if the crystallization kinetics of the phase has been described in some papers [2] [3], there is a lack of information regarding the crystallization process at low and high undercooling and the final phase composition as a function of temperature. This can be due to the very demanding cooling rates needed to allow isothermal crystallization from the melt at very low temperatures. ...
... The resulting values of the parameters for heterogeneous nucleation density are listed in Table 1. For the homogeneous nucleation rate, only e 2 has been used as fitting parameter while the h f and T are taken from literature [35] [3]. ...
Article
The crystallization of Polyamide 12 (PA12) has been investigated using a new experimental setup which allows in-situ synchrotron Wide Angle X-ray Diffraction (WAXD) during Flash-DSC measurements. The experimental results are used to parameterize and validate a new numerical model to quantify the quiescent crystallization kinetics, under isothermal conditions, of the three important crystal structures of PA12, i.e. the, γ− α′− and mesomorphic phase. The experimental approach is based on nucleation and growth after quenching the material from the melt to an isothermal temperature, described by the Schneider rate equations and the Kolmogorov-Avrami expression for the space filling. The experimental overall crystallization rate, expressed in terms of the crystallization half-time, as well as the phase composition, are well captured by the model over a wide range of temperatures, i.e. between the glass transition and the melting temperature. It is shown that at temperature below and above 100°C different nucleation mechanisms are dominant causing the bimodal dependence of the crystallization rate. This work forms the basis for a full model for non-isothermal conditions for which transitions between different phase have to be taken into account.
... A literature review yields several reports that describe modeling for PA12 [9][10][11][12]. However, the literature on the comparison among different non-isothermal models has not been presented. ...
Article
Full-text available
Selective laser sintering (SLS) of thermoplastic materials is an additive manufacturing process that overcomes the boundary between prototype construction and functional components. This technique also meets the requirements of traditional and established production processes. Crystallization behavior is one of the most critical properties during the cooling process and needs to be fully understood. Due to the huge influence of crystallization on the mechanical and thermal properties, it is important to investigate this process more closely. A commercial SLS polyamide (PA12) powder was measured with differential scanning calorimetry (DSC) to model a wider temperature range. To model isothermal crystallization between 160 and 168 °C, the Avrami model was used to determine the degree of crystallization. For non-isothermal crystallization between 0.2 and 20 K/min, different models were compared including the Ozawa, Jeziory, and Nakamura equations.
... The signal for DF-PA is much broader and pronounced compared to Orgasol-IS. Due to the formation of the α-crystalline-structure during the precipitation of DF-PA, the area for the XRD-signal is much wider and indicates a higher crystallinity [24]. ...
... The signal for DF-PA is much broader and pronounced compared to Orgasol-IS. Due to the formation of the α-crystalline-structure during the precipitation of DF-PA, the area for the XRD-signal is much wider and indicates a higher crystallinity [24]. The distinct melting behaviors can be related to different crystalline structures, since an additional tempering step is carried out to obtain higher values for crystallization enthalpy (ΔHm) compared to "Standard-PA12" (see Table 1). ...
Article
Full-text available
Different features of polymer powders influence the process of laser sintering (LS) and the properties of LS-parts to a great extent. This study investigates important aspects of the “powder/process/part”-property relationships by comparing two polyamide 12 (PA12) powders commercially available for LS, with pronounced powder characteristic differences (Duraform® PA and Orgasol® Invent Smooth). Due to the fact that the primary influence factor on polymer behaviour, the chemical structure of the polymer chain, is identical in this case, the impacts resulting from powder distribution, particle shape, thermal behaviour, and crystalline and molecular structure, can be studied in detail. It was shown that although both systems are PA12, completely different processing conditions must be applied to accomplish high-resolution parts. The reason for this was discovered by the different thermal behaviour based on the powder production and the resulting crystalline structure. Moreover, the parts built from Orgasol® Invent Smooth unveil mechanical properties with pronounced anisotropy, caused from the high melt viscosity and termination of polymer chains. Further differences are seen in relation to the powder characteristics and other significant correlations could be revealed. For example, the study demonstrated how the particle morphology and shape impact the surface roughness of the parts.
... The kinetic parameters used to monitor the effect of Cr 2 O 3 nanoparticles on PA12 matrix crystallization were calculated from the partial integration of the DSC crystallization exotherms, with experimental error of around 1%. For the estimation of crystallinity of the samples, a value of 95 Jg-1 for the melting enthalpy of 100% crystalline PA12 was used [12]. The degree of crystallinity, v, was calculated from the areas of the corresponding DSC melting peaks according to: ...
Article
Polymer nanocomposites based on PA12 filled with different loading (0.1−10 wt%) of nanosized (average grain size of about 1−5 nm) chromium(III) oxide were prepared by in situ polymerization, and their mechanical and thermal properties were investigated. A homogeneous dispersion of Cr2O3 nanoparticles in the polyamide matrix was achieved at the level of isolated particles as a result of amphoteric nature of chromium (III) oxide. However, it was found that the tensile modulus and yield stress of the nanocomposites decrease. The reason for this is a change in the crystallization type of PA12due to the Cr2O3 nanoparticles (from α to γ type of crystal phase). The crystallization kinetic of PA12 matrix has been successfully studied and described using a two-parameter Avrami model for isothermal crystallization. It is shown that nanosized chromium(III) oxide dioes not change the nucleation mechanism. The addition of Cr2O3 nanoparticles also does not have a significant effect on crystallization rate and energetic and kinetic parameters of crystallization. The presence of nanofillers in PA12 matrix decreases their crystallinity from 63 to 51%. Hence, the mechanical properties of nanocomposites can be influenced not only by the type and dispersion of nanofillers but also by their effect on the crystal structure and degree of crystallinity of the polymer matrix. POLYM. COMPOS., 2015. © 2015 Society of Plastics Engineers
... For the first two processes clearly the objective is the design of a next generation of predictive software to be used in industry to support the development of process windows. Once the thermal history is fully known Avrami-type of crystallinity models are used to predict the degree of crystallinity in different areas of the part [2]. In the simulation of laser cutting the focus is on the understanding of the different mechanisms, e.g. ...
Conference Paper
Full-text available
Thermoplastic composites feature the advantage of melting and shaping. The material properties during processing and the final product properties are to a large extent determined by the thermal history of the material. The approach in the FP7-project FibreChain for process chain modeling of thermoplastic composites is presented.
... Les points expérimentaux de G en fonction de la température sont bien décrits par le modèle de Lauritzen-Hoffman incluant ces valeurs (cf. Figure B.28 (b)). Il semble que dans l'intervalle de température choisi il n'y ait pas de transition de régime, ce qui est conforme aux observations de Neil et al [Neil 2008] et Plummer et al [Plummer 2001]. Pour modéliser la cristallisation selon la théorie d'Avrami, la constante d'Avrami a tout d'abord été recalculée grâce à l'Equation B.10, par l'intermédiaire du temps de demicristallisation pour le cas d'une germination sporadique, c'est-à-dire pour une valeur de l'exposant d'Avrami n égal à 4. Cette valeur va nous permettre de relier le nombre de germes, la fréquence d'activation, la vitesse de croissance des sphérolites et la constante d'Avrami k comme l'exprime l'Equation B.9. La Figure B.29 montre qu'il existe une dépendance linéaire de ln(k) en fonction de la température sur la zone de température considérée. ...
Article
Full-text available
Additive processing technologies are aimed at manufacturing parts directly from a computer-aided design (CAD) file, without the need for tooling. Therefore flexibility of production increases and manufacturing of small to mid-size series of very complex or even customized parts becomes possible within reduced development time and expenses. Because of the good mechanical properties obtained in the parts, the most commonly used among additive technologies for polymers is laser sintering (LS). The objective of this work is to contribute to a better understanding of the different physical mechanisms involved during laser sintering of polyamide 12 powders. Many operating variables impact the laser sintering process. Especially, the energy supplied to the powder with the laser beam depends on its power, its displacement velocity and the scan spacing. Moreover, the polymer material undergoes a quite severe thermal treatment : before its sintering, the powder is preheated, then in the build tank the sintered parts and the un-sintered surrounding powder remain until the end of the job at elevated temperatures. This thermal history induces ageing, which modifies some powder features and hinders its future reuse. The influence of the parameters mentioned above on the part microstructure and mechanical properties was investigated. Moreover the use of different polyamide 12 powders enabled to identify the key material characteristics towards the physical processes involved in LS and towards the final properties of parts. The laser sintering of semi-crystalline polymers is governed by several fundamental mechanisms: melting of particles, interdiffusion of macromolecular chains at interfaces, coalescence of molten particles, then densification and finally crystallisation. The study and modelling of crystallisation were carried out with one of the PA12 powders used in the first part of this work. From this modelling, the time during which the polymer remains in the molten state during the process was estimated. Next, a rheological analysis made within the framework of linear viscoelasticity of polymer melts allowed to compute the interdiffusion time of the macromolecular chains. Moreover, the coalescence process of molten particles was observed at different temperatures and modeled. The characteristic times thus estimated for these physical processes were opposed to the time during which the polymer remains in the molten state and confirmed the good consolidation obtained by laser sintering of polyamide 12. In conclusion this work contributes to understand the different physico-chemical mechanisms involved during polymer laser sintering by specifying the relations between powder parameters, process variables and final properties of parts. Many recommendations for the optimisation of powder properties can be derived from this work for the purpose of developing new polymeric materials adapted to this process.
... By using thermal measurements via differential scanning calorimetry (DSC), the glass transition temperature (T g) and the thermal stabilities of the polymers can be characterised (Hoekstra et al 1995;Khanna et al 1995;Gajdoš et al 2001). The crystallinity also defines the physical characteristics of the material such as the stability in different processes (Bonnet et al 1998;Demertzis & Franz 1998;Fuchs et al 1998;Plummer et al 2001;Zhu & Yan 2001). ...
Article
Full-text available
This work is an experimental study of the differential scanning calorimetry characterisation of polymer materials used in food packaging materials, such as polypropylene (0.03 mm), polyethylene (0.1 and 0.03 mm), poly(D-(-)-Β-hydroxybutyrate) (powder), two-layered polypropylene (0.064 mm), and two-layered polypropylene with poly-vinylidene-chloride (0.012/0.021). The polymer stability was checked by simulation of conditions during food preparation in microwave ovens, sterilisation or rapid freezing. The materials were tested in the temperature range from 40 to 200‡C at different scan rates from 2 to 30°C min−1 during heating or cooling. The enthalpies show a high correlation coefficient (0.964) with scan rate. All samples undergo phase change in the temperature range from 107 to 173°C during heating and enthalpies are in the range from 31.8 to 71.1Jg−1. Upon subsequent cooling from 200°C, the temperature range of phase changes is shifted to lower temperatures from 86 to 102°C with enthalpies ranging from 30.4 to 57.8 J g−1. Experiments with exposure of polymers to microwave radiation and freezing prove that the phase change considering the temperature range is very similar in all experiments.