• No results found

A new engineering approach to predict the long-term hydrostatic strength of unplasticized poly(vinyl chloride) pipes

N/A
N/A
Protected

Academic year: 2021

Share "A new engineering approach to predict the long-term hydrostatic strength of unplasticized poly(vinyl chloride) pipes"

Copied!
11
0
0

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

Hele tekst

(1)

A NEW ENGINEERING APPROACH TO PREDICT THE

LONG-TERM HYDROSTATIC STRENGTH OF

UNPLASTICIZED POLY(VINYL CHLORIDE) PIPES

H.A. Visser†*, M. Wolters†, T.C. Bor†, T.A.P. Engels‡, L.E. Govaert‡ †

University of Twente, P.O. Box 217, 7500 AE, Enschede, the Netherlands ‡

Eindhoven University of Technology, P.O. Box 513, 5600 MB, Eindhoven, the Netherlands

*

Corresponding author: h.a.visser@ctw.utwente.nl

ABSTRACT

Extruded polymer pipes are qualified using pressurized pipe tests. With these tests the long-term hydrostatic strength is determined by subjecting the pipes to an internal pressure, while measuring the time-to-failure. Although these tests can be accelerated (at higher temperatures), they remain time consuming and require a spacious experimental setup. To circumvent this costly method a model based approach is proposed by which the long-term hydrostatic strength is predicted. Using short term measurements, the input parameters for this approach can be determined. In this engineering approach the effects of physical aging are included. The approach is capable to quantitatively predict the (long-term) failure time of pipe sections under internal pressure.

Unplasticized Poly(Vinyl Chloride), pipe, lifetime assessment, physical aging

1 INTRODUCTION

A well-known structural application of polymers can be found in the water and gas distribution industry where polymer pipes are used extensively for transportation of fluids. The long-term hydrostatic strength of these pipes is currently determined using internally pressurized pipe tests (also referred to as burst tests). ISO 1167 describes the procedure for carrying out these tests, where sections of a pipe are subjected to predefined internal pressures and several temperatures, while the time-to-failure is measured. The long-term hydrostatic strength can be extrapolated according to ISO 9080 from the obtained data. Unfortunately, burst tests still last up to typically a year. In combination with the spacious experimental setup required, this step in the qualification process is costly. Therefore, it has been a scientific challenge since several decades to predict the long-term hydrostatic strength based on short term measurements [1]. It still is a challenge, proven by the fact that recently several studies were presented on this subject, like [2-5]. In a previous paper [6] we presented an approach that is able to quantitatively predict the lifetime of a pressurized unplasticized poly(vinyl chloride) (uPVC) pipe, based on only one short-term measurement which gives a reference point. In this paper this approach will be extended to include phenomena related to physical aging in the predictions.

(2)

2 THEORY Our approach is based upon the following hypothesis:

1. During burst tests three types of failure can be encountered: ductile tearing, hairline cracking and brittle fracture. Despite the obvious macroscopic differences, all failure modes are initiated by (local) plastic deformation.

2. Upon applying an internal pressure, plastic deformation will accumulate in a plastic pipe. The rate of accumulation follows an Eyring type reaction rate relation [7].

3. Failure occurs when the accumulated plastic strain equals critical plastic strain, where the polymer enters its softening region. This critical plastic strain is constant for a wide range of applied stresses and temperatures.

4. The effects of physical aging can be incorporated in the model by including aging kinetics in the Eyring type equation.

These four steps in our hypothesis will be described in more detail in the following sections.

2.1 Failure Initiation

During burst tests uPVC pipes will either fail after large macroscopic plastic deformation (ductile tearing), small macroscopic deformation followed by a crack in longitudinal direction (hairline cracking), or after no visible macroscopic plastic deformation (brittle fracture). Niklas et al. [8] observed all of these three failure modes. Despite the distinct differences in macroscopic deformation behaviour, the failure kinetics appeared to be identical. Niklas et al. noted this indicates all three failure modes had one underlying failure mechanism. Their hypothesis is supported by Kramer [9] who states that plastic deformation precedes crazing, which leads to brittle fracture. Thus, although no plastic deformation is observed on a macroscopic scale, brittle fracture is initiated by local plastic deformation. Therefore, first the focus will be on the kinetics of this plastic deformation, to eventually predict the time-to-failure.

2.2 Deformation Kinetics

There have been several successful attempts to describe the deformation kinetics (mostly the yield stress versus strain rate and temperature as measured in uni-axial tension) of several polymers [10-15]. The yield behaviour of poly(carbonate) (PC) can be described with an Eyring type relation [7] over a relatively wide temperature range [10,11]. However, it is well known that glassy polymers show a secondary transition for deformation at higher rates, or at lower temperatures. For most glassy polymers the temperatures and deformation rates where only the primary transition contributes to the yield behaviour is limited to a high temperatures or low strain rates, thus to long-term applications. In case two mechanisms are active, a Ree-Eyring relation [16] can be employed to describe the deformation kinetics, as has been done for uPVC [11] and other polymers, e.g. [12-15]. Although the deformation behaviour of uPVC has a distinct contribution of the secondary transition at lower temperatures and higher strain rates, long-term creep failure is dominated by the contribution of the primary transition, as has been shown for PC by Klompen et al. [17]. Therefore, in this study only the primary transition will be taken into account for uPVC.

(3)

In this specific case the deformation kinetics of a pipe subjected to an internal pressure is of interest. It is already shown by various authors that the steady state reached under constant stress conditions, equals that reached under constant strain rate [12,18-21]. In other words: the plastic deformation behaviour during creep tests can be characterized using constant strain rate tests. Furthermore, Ward [22] proposed a pressure-modified Eyring flow relation in which the influence of the hydrostatic pressure is quantified. With this addition it is possible to describe the deformation kinetics for polymers loaded under various multi-axial stresses with one parameter set only [10,23,24]. This enables us to predict the deformation kinetics of a pressurized pipe, based upon a characterization using constant strain rate tests as measured in uni-axial tension.

The pressure modified Eyring type equation used here, relates the equivalent strain rate (γ& ) with the applied equivalent shear stress (τ ) and the hydrostatic pressure (p).

(

)

 −       ∆− = kT V kT pV RT U p T * *

0exp exp sinh

,

,τ γ µ τ

γ& & (1)

This equation consists of four terms: a pre-exponential factor (γ& ) which changes with 0 the thermodynamic state of the material, an Arrhenius type relation for the temperature activation with activation energy (ΔU), a term for the stress activation with activation volume (V*) and a term that incorporates the influence of the hydrostatic pressure with the pressure dependence (μ). R and k are the Gas and the Boltzmann constants respectively. In uni-axial tension the equivalent strain rate (γ& ), the equivalent shear stress (τ ) and hydrostatic pressure (p) are given by ε&3 ,

3 σ and σ 3 1 − respectively. 2.3 Failure Criterion

Klompen et al. [25] were able to quantitatively predict the long-term failure of PC tensile bars using a 3D finite strain approach. Their model identified unstable deformation, leading to failure, based on the intrinsic behaviour of the material, without a pre-described failure criterion. The instabilities occur just after yielding, when the material enters its softening region. The softening function Klompen et al. used was independent of the applied stress or temperature, thus the polymer entered its softening region at a constant value for the plastic strain. This suggests a critical plastic strain criterion is suitable for quantitative failure time predictions, which is supported by the positive experiences in the past when using such a criterion [18,19,26]. The time-to-failure (tf) for a geometry that deforms at a constant rate (as a result of a constant

applied stress at constant temperature) can be calculated with:

(

T p

)

t cr f , ,τ γ γ & = , (2)

where γ is the critical equivalent plastic strain. By combining this equation with eq. cr (1) and using the approximation that sinh(x)≈0.5exp(x) for x>>1, a relation for the time-to-failure for a uni-axial creep test (at constant engineering load) can be derived:

( )

(

)

    + − ∆ = µ σ γ γ σ kT V RT U T t cr f 3 3 exp 2 , * 0 & . (3)

(4)

2.4 Incorporating Physical Aging

Struik [27] showed that physical aging has a significant influence on the deformation kinetics of glassy polymers. So far physical aging is not accounted for, as the pre-exponential factor (γ& ) is not a function of time yet. Here, an approach identical to that 0

of Klompen et al. [28] will be used to incorporate physical aging effects in the deformation kinetics. The pre-exponential factor is replaced by the following relation:

(

T, ,t

)

0,rejexp

(

Sa

(

T, ,t

)

)

0 τ γ τ

γ& = & − . (4)

Where γ&0,rej is a material constant that is related to the rejuvenated state of the polymer and Sa characterizes the thermodynamic state of the polymer. The Sa and the effective

aging time (teff) are related as follows:

(

)

(

eff

(

)

a

)

a T t c c t T t t

S ,τ, = 0 + 1log ,τ, + , (5) with c0 and c1 constants and ta the initial age of the polymer. It is well known that aging

is accelerated at both elevated temperatures and stresses, as increasing temperature and stress both increase the mobility of the polymer chains. To take this into account the effective aging time is introduced, where the aging time at a certain temperature and under a certain stress is translated towards the aging time at a reference temperature and stress. The reference temperature (Tref) and stress are taken as 25˚C and 0 MPa

respectively throughout this work. The following equation is used to convert the aging time (t) into effective aging time:

(

)

=

( )

( )

(

( ) ( )

)

t T eff t t T a t T a t t T t 0 ' ', ' ' d , , τ τ σ , (6) with

( )

( )

( )

               − ∆ = ref a T T t T R U t T a exp 1 1 (7) and

( ) ( )

(

,

) ( )( )

sinh

( )

( )

1 −           = t RT t t RT t t t T a τ τ υa τ υa σ . (8)

The aging activation energy (ΔUa) and the specific aging activation volume (υ ) are the a

parameters that determine the influence of temperature and stress on aging. Equation (2) does not hold for experiments in which the deformation kinetics changes as a result of physical aging (which is the case if teff>ta). The criterion is therefore changed into:

(

T p t

)

t f t cr , , , d 0 τ γ γ =

& . (9)

The theory as presented in this section will be validated using burst test data as measured by Niklas et al. [8]. Before the validation step, the procedure to characterize the deformation and aging kinetics will be described briefly. The experimental methods used for this characterization step will be described in the next section.

(5)

3 EXPERIMENTAL 3.1 Material and Sample Preparation

The uPVC samples were taken out of an excavated uPVC gas distribution pipe (Ø160 mm) that had been in service for several decades. With a bandsaw a section of 70 mm was cut from the pipe material. This section was sawed in axial direction. These parts were then pressed into flat plates in a press at 100°C, followed by fast cooling in a cold press at 20°C. Tensile bars with a parallel gauge section of approximately 30×5×4 mm3 were milled from the plate material. Three sets of samples were used. The first set was used to characterize the deformation kinetics of uPVC. To prevent a significant contribution of physical aging to the deformation kinetics during these experiments, this set of samples was given an additional heat treatment after milling. The samples were aged for 5·105 s at 60°C. The other two sets were used for characterizing the temperature and stress induced aging kinetics respectively. These samples received different, specified heat or stress treatments after milling.

3.2 Experimental Setup

All uni-axial tensile and creep measurements were carried out on a MTS Elastomer Testing System 810 equipped with a 25 kN force cell and a temperature controlled chamber. The tests for characterizing the deformation kinetics were conducted at three different temperatures (20°C, 40°C and 60°C). All tests to characterize the aging kinetics were conducted at a temperature of 25°C and an engineering strain rate of 10-3 s-1. The engineering stresses were calculated using the average of the cross sectional surface areas as measured at three locations in the gauge section. Tensile experiments were carried out at a constant crosshead speed, thus at a constant engineering strain rate. The creep tests were conducted at constant load, thus at constant engineering stress. To avoid dynamic effects the load is applied within 3 seconds. All stresses and strains reported in this paper are engineering values.

4 CHARACTERIZATION PROCEDURE

The goal of this work is to study whether it is possible to quantitatively predict long-term failure of uPVC pipes subjected to an internal pressure, using the hypothesis as presented in section 2. Before the validation of the approach is presented, the procedure to characterize the deformation and aging kinetics is described.

4.1 Characterizing Deformation Kinetics

The deformation kinetics are determined by material parameters μ, ΔU and V* and the thermodynamic state of the material, which is represented in γ& and is considered 0 constant in this part of the characterization procedure. The pressure dependence (μ) was determined using yield data for tensile bars under a superimposed hydrostatic pressure as measured by Yuan et al. [29]. The pressure dependence can be derived from the slope of the yield stress versus the superimposed hydrostatic pressure (po), by rewriting eq. (1)

(6)

3 3 d d + = µ µ σ o p . (10)

The engineering yield data versus the superimposed pressure as measured by Yuan et al. [29] is shown in figure 1. A least squares fit on this data using eq. (10) resulted in

μ=0.14. A line with a slope corresponding to this value is also shown in figure 1.

The tensile yield stress was measured at several strain rates and at three different temperatures (20°C, 40°C and 60°C) to determine the remaining parameters that characterize the deformation kinetics (ΔU, V* and γ& ). At each combination of strain 0

rate and temperature three tests were carried out. The experimental results are shown in figure 2. Very little scatter was found between the individual yield stress measurements and, as expected, an increase in yield stress was found at increasing strain rates and decreasing temperatures. The parameters were determined by applying a least squares fit on all yield data, using eq. (1) and the previously determined value for μ. The resulting values are: V*=3.42 nm3, ΔU=297 kJ·mol−1 and γ& =2.9·100 36 s−1. It is Figure 1: Yield stress as measured in tension with a superimposed hydrostatic pressure.

The solid line is a best fit using eq. (10) resulting in μ=0.14. Data taken from [29].

Figure 2: Yield stress versus strain rate at three temperatures. The solid lines represent the best fit using eq. (1) and ΔU,

V* and γ& as fitting parameters. 0

Figure 3: Time-to-failure for tensile creep experiments at three temperatures. The solid lines represent the best fit using the critical strain as the fit parameter in eq. (3).

(7)

emphasized that this value for γ& is unique for this set of samples, and will differ for 0 samples with a different thermo-mechanical history and changes as the material ages. The solid lines in figure 2 represent the best fit using eq. (1) and show that the tensile yield behaviour of uPVC can be accurately described with the pressure modified Eyring relation. The values for the activation volume and the activation energy are close to the values found by Bauwens-Crowet et al. [11] for uPVC (3.22 nm3 and 295 kJ·mol−1 respectively).

4.2 Determining the Critical Equivalent Plastic Strain

From the postulated theory it is expected that the time-to-failure in tensile creep tests can be predicted using eq. (3), thus assuming a constant value for the equivalent critical plastic strain (γ ). Figure 3 shows the time-to-failure as measured for various applied cr stresses at three temperature levels. The solid lines represent the calculations using eq. (3), the previously found material parameters and a fitted critical equivalent strain with a value of 1.5%. The employed relation appears to take the influence of both temperature and stress appropriately into account, describing the experimental results accurately.

4.3 Characterizing Physical Aging Kinetics

The aging kinetics are captured with the parameters γ&0,rej, c0, c1, ΔUa and υ . This a

paper will not cover a complete description how these parameters are determined, but focuses on the latter two. A concise description on how to determine the other parameters can be found in [28].

The temperature-activated aging was characterized using tensile yield data of samples that received heat treatments at different temperatures for various aging times, as shown in figure 4. The activation energy (ΔUa) was computed from the shift towards a

mastercurve (shown in figure 5) using equations (6) and (7) and was found to be 113 kJ∙mol-1. The location of the mastercurve is determined by the parameters γ&0,rej, c0 and

c1, which were found using the intrinsic mechanical behaviour of uPVC (not described

here) and are 2.4∙1041 s-1, -6.24 and 2.20 respectively. In figure 5 the mastercurve for

Figure 4: Yield stress measured at a strain rate of 10-3 s-1 and 25°C for samples that

were aged at four temperatures.

Figure 5: Data from figure 4 shifted towards the master curve. The resulting best fit is obtained with ΔUa=113 kJ∙mol-1.

(8)

samples with an initial age of 3.5∙105 s tested at a temperature of 25°C and a strain rate of 10-3 s-1 is shown. Additionally, the shifted yield data is shown, which reveals that the mastercurve describes the yield behaviour very well.

The stress induced aging kinetics were determined in a similar way. In this case, the yield data of samples which were loaded at a tensile stress of 25 MPa prior to testing was used (see figure 6). The activation volume followed from the shift towards the mastercurve in figure 7 (ta=2.1∙106 s), using equations (6) and (8) a value of 1.03∙10-3

m3∙mol-1 was found. This value should be used with caution as it is determined using data at one aging stress level only. Further experimental research at a range of aging stress levels is needed to obtain a more reliable value for this parameter. For now, the obtained value is a good approximation.

5 VALIDATION: PREDICTING BURST TEST DATA

To test the predictive capacity of the presented approach, burst test data of Niklas et al. [8] on uPVC pipes will be used. They carried out burst tests at temperatures of 20°C, 40°C and 60°C. Their results are reproduced in figure 8, in which the failure modes are indicated: ductile tearing (unfilled markers), hairline cracking (grey markers) and brittle fracture (black markers). Although there are distinct differences between the three types of failure, the failure kinetics appear to be identical.

The stress state in the pipe during a burst test is given by the stress in the circumferential (referred to as hoop stress, σh), longitudinal (σl), and the radial direction

(σr). When a constant internal pressure (pi) is applied these stresses can be calculated

using Barlow’s formula for thin walled pressure vessels:

(

)

w i w h t p t D 2 − = σ ,

(

)

w i l t p t D 4 − = σ and σr =0. (11)

with D, the outer diameter of the pipe and tw, the wall thickness. Combining equations

(1), (2) and (12), table 1 and using the approximation that sinh(x)≈0.5exp(x) for x>>1, results in the following relation for the time-to-failure for pipes with an internal pressure

pi:

Figure 6: Yield stress measured at a strain rate of 10-3 s-1 and 25°C for samples that were aged at an applied tensile stress of 25

MPa.

Figure 7: Data from figure 4 shifted towards the master curve. A least squares

fit with eq. (6,8) results in υ =1.03∙10a -3 m3∙mol-1.

(9)

(

)

= ∆ −

(

)(

)

iw w cr i f p kT t V t D RT U p T t 4 1 exp 2 , * 0 µ γ γ . (12)

In this relation the influence of physical aging is neglected. To incorporate aging effects in the time-to-failure predictions equations (1), (4-6) and (9) were combined. This resulted in the following relation between the time-to-failure for samples under constant stress and temperature:

(

)

( ) ( )

(

)

( ) ( )

                −               +     + − ∆     − = − − a c c a T cr T f t t c kT pV RT U kT V T a T a c T a T a p T t 2 2 1 1 1 0 * * 2 exp sinh , 1 , , , µ τ τ γ τ τ σ σ K .(13)

With c1 =c2ln

( )

10 , and for pipe under internal pressure:

(

)

i

w w p t t d 4 − = τ and p=−τ .

Before the predictions can be calculated, the initial age of the samples used by Niklas et al. should be known. Therefore a reference point is used to calculate γ& in eq. (12) and 0

ta in eq. (13), resulting in a value of 1.4·1036 s−1 and 2.1∙108 s respectively. With these

values a yield stress of 62 MPa is predicted for a tensile test at a strain rat of 10−3 s−1 at 20°C, which is a realistic value for uPVC.

The solid lines in figure 8 represent the predictions using eq. (12), the material parameters as derived for uPVC during the current research and the reference point indicated in the figure. Not only the slope of the data points is predicted accurately, but also the influence of the temperature agrees very well with the experimental values. These results indicate that the presented approach is capable of predicting failure of a

Figure 8: Failure time for uPVC pipes under internal pressure, taken from [8]. The unfilled, grey and black markers represent ductile tearing, hairline cracking and brittle fracture respectively. The solid lines are predictions using eq. (12) (without aging) and

(10)

uPVC structure on which a 3D load is applied. Moreover, it proves its applicability to samples with a different thermodynamic state, as the pipes Niklas et al. used were “older” than those used for the characterization of uPVC during this study.

The dashed lines in figure 8 represent the predictions using eq. (13), where the influence of physical aging is taken into account. The tests conducted at 60°C show the most distinct influence of physical aging, revealing an endurance limit. This is mainly induced by the elevated temperature, whereas the measurements at 20°C are aged mainly due to the applied stress. The predictions clearly suggest the presence of an endurance limit. However, the experimental data of Niklas et al. does not show a significant influence of physical aging. Further validation of the aging kinetics will therefore be the scope of future work.

6 CONCLUSIONS

A new engineering approach for predicting the time to failure of loaded glassy polymers has been presented. The approach is founded on an hypothesis that consists of four parts:

− Failure is initiated by plastic deformation, irrespective of the failure mode.

− The accumulation of plastic deformation can be predicted using a pressure modified Eyring relation.

− Failure occurs when the accumulated plastic strain is equal to a critical strain, where the polymer enters its softening region leading to localized plastic deformation.

− The influence of physical aging can be incorporated by including an extra term in the pressure modified Eyring relation.

With this hypothesis the time-to-failure of a pressurized uPVC pipe can be predicted quantitatively, using only short term experiments. By incorporating the influence of physical aging an endurance limit is observed in the predictions. Further research is needed to validate the aging kinetics.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the financial support provided by four Dutch gas distribution network service providers: Cogas Netbeheer, Continuon, Eneco Netbeheer and Essent Netwerk.

REFERENCES

[1] R.W. Lang et al. (1997) Angew. Makromol. Chem., 247:131.145 [2] N. Brown (2007) Polym. Eng. Sci., 47:477–480

[3] G. Pinter et al. (2007) Monatsh. Chem., 138:347–357

[4] S. Castagnet et al. (2004) Consequences of annealing during long-term pressure tests on PolyVinylidene Fluoride pipes, In: Proceedings of Plastic Pipes XII, Baveno, Italy

[5] G. Castiglioni et al. (2004) Prediction of ductile failure in u-pvc pipes from creep tests on specimens, In: Proceedings of Plastic Pipes XII, Baveno, Italy

[6] H.A. Visser et al. (2007) A new engineering approach to predict the hydrostatic strength of uPVC pipes, In: Proceedings of PPS regional meeting E/A, Gothenburg, Sweden

(11)

[7] H. Eyring (1936) J. Chem. Phys., 4:283–291 [8] H. Niklas et al. (1963) Kunststoffe, 53:886–891 [9] E.J. Kramer (1983) Adv. Pol. Sci., 52/53:1–56 [10] R.A. Duckett et al. (1978) Brit. Polym. J.,10:11–16

[11] C. Bauwens-Crowet et al. (1969) J. Polym. Sci. Pol. Phys., 7:735–7421 [12] J.A. Roetling (1965) Polymer, 6:311–317

[13] J.A. Roetling (1965) Polymer, 6:615–619 [14] J.A. Roetling (1966) Polymer, 7:303–306

[15] L.E.Govaert et al. (2000) Polymer, 41:1959–1962 [16] T. Ree et al. (1955) J. Appl. Phys., 26:794-800

[17] E.T.J. Klompen et al. (1999) Mech. Time-Dependent Mater., 3:49–69 [18] M.J. Mindel et al. (1973) J. Mater. Sci., 8:863–870

[19] C. Bauwens-Crowet et al. (1974) J. Mater. Sci. Lett., 4:1197–1201 [20] R.M. Ogorkiewicz et al. (1967) Brit. Plas., 40:126–128

[21] I.M. Ward et al. (1984) J. Polym. Sci. Pol. Phys., 22:561–575 [22] I.M. Ward (1971) J. Mater. Sci., 6:1397–1417

[23] L.E. Govaert et al. (2000) Mech. Time-Dependent Mater., 4:275–291 [24] A.R. Ragab (2001) J. Appl. Polym. Sci., 81:991–999

[25] E.T.J. Klompen (2005) Macromolecules, 38:7009–7017 [26] D.J. Matz (1972) J. Polym. Sci. Pol. Phys., 10:1017–1930

[27] L.C.E. Struik (1978) Physical aging in amorphous polymers and other materials Elsevier, Amsterdam, the Netherlands

[28] E.T.J. Klompen et al. (2005) Macromolecules, 38:6997–7008 [29] J. Yuan et al. (1983) J. Mat. Sci., 18:3063–3071

Referenties

GERELATEERDE DOCUMENTEN

For this research perceived usefulness and perceived ease of use are used as starting points from which experiences and feelings of employees on the process of

Mastery Experiences en PCK-in praktijk zijn weliswaar kenmerken uit de professionele muzikale achtergrond van leerkrachten, maar vermoed wordt dat ze een mediërende

Tot slot, doordat cultuur in deze thesis geclassificeerd is in de dimensies van Hofstede, kan vergeleken worden welke cultuurkenmerken (nationale cultuur, landspecifieke

The bollworm complex consisting of the African bollworm Helicoverpa armigera (Hübner) (Noctuidae), red bollworm, Diparopsis castanea Hampson (Noctuidae), and spiny

Rapporten van het archeologisch onderzoeksbureau All-Archeo bvba 321 Aard onderzoek: opgraving Vergunningsnummer: 2016/244 Naam aanvrager: Natasja Reyns Naam site: Hoogstraten

4 § 2 van het Decreet houdende Bescherming van het Archeologisch Patrimonium van 30 juni 1993 (B.S. 7.6.2006) zijn de eigenaar en de gebruiker ertoe gehouden de

Het management van produktieprocessen kan men onderscheiden naar de beheersing van het primaire proces en de beheersing van een aantal ondersteunende processen.

This resulted in a second transition, where outsourced housecleaning service companies transformed this personal maternalistic employment relationship into a triangular