• No results found

Flames in channels with cold walls: acceleration versus extinction

N/A
N/A
Protected

Academic year: 2022

Share "Flames in channels with cold walls: acceleration versus extinction"

Copied!
11
0
0

Loading.... (view fulltext now)

Full text

(1)

http://www.diva-portal.org

Postprint

This is the accepted version of a paper presented at Ninth Mediterranean Combustion Symposium, 7-11 June 2015, Rhodes, Greece.

Citation for the original published paper:

Dion, C., Demirgök, B., Akkerman, V., Valiev, D., Bychkov, V. (2015) Flames in channels with cold walls: acceleration versus extinction.

In:

N.B. When citing this work, cite the original published paper.

Permanent link to this version:

http://urn.kb.se/resolve?urn=urn:nbn:se:umu:diva-107878

(2)

MCS 9 Rhodes, Greece, June 7-11, 2015

FLAMES IN CHANNELS WITH COLD WALLS:

ACCELERATION VERSUS EXTINCTION

C. M. Dion*, B. Demirgok**, V. Akkerman**, D. Valiev*** and V. Bychkov*

Vitaliy.Bychkov@physics.umu.se

*Dept. Physics, Umeå University 90187 Umeå, Sweden

**Dept. Mechanical and aerospace Engineering, West Virginia University, 26506 Morganton, USA

***Dept. Applied Physics, Umeå University 90187 Umeå, Sweden

Abstract

The present work considers the problem of premixed flame front acceleration in micro- channels with smooth cold non-slip walls in the context of the deflagration-to-detonation transition; the flame accelerates from the closed channel end to the open one. Recently, a number of theoretical and computational papers have demonstrated the possibility of powerful flame acceleration for micro-channels with adiabatic walls. In contrast to the previous studies, here we investigate the case of flame propagation in channels with isothermal cold walls. The problem is solved by using direct numerical simulations of the complete set of the Navier- Stokes combustion equations. We obtain flame extinction for narrow channels due to heat loss to the walls. However, for sufficiently wide channels, flame acceleration is found even for the conditions of cold walls in spite of the heat loss. Specifically, the flame accelerates in the linear regime in that case. While this acceleration regime is quite different from the exponential acceleration predicted theoretically and obtained computationally for the adiabatic channels, it is consistent with the previous experimental observations, which inevitably involve thermal losses to the walls. In this particular work, we focus on the effect of the Reynolds number of the flow on the manner of the flame acceleration.

Introduction

Deflagration-to-detonation transition (DDT) is one of the most important and fascinating combustion phenomena with wide range of applications from pulse-detonation engines to safety issues such as the prevention of mining accidents [1-3]. The reason and the key element of DDT is spontaneous flame acceleration in tubes and channels from a low laminar flame speed to nearly sonic values, which implies ultra-sonic propagation in the reference frame of the tubes walls by the end of the process [4-6]. Two major mechanisms of flame acceleration in channels have been elucidated: the Shelkin mechanism in the case of smooth walls [7,8]

and the mechanism of ultra-fast flame acceleration in channels with obstacles [9]; the present work focuses on a geometry with smooth walls.

The most typical geometry of DDT in experiments and energy-production devices corresponds to a relatively long channel, with the flame propagating from the closed channel end to the open one. According to the Shelkin mechanism, a flame front in channels with smooth walls accelerates because of the thermal expansion of the burning gas and the non-slip boundary conditions at the walls. Expansion of the burning gas produces a flow of the fuel mixture, which becomes non-uniform due to the non-slip boundary conditions. The non- uniform flow makes the flame front curved, which increases the burning rate and creates a positive feedback between the flame and the flow, hence leading to the flame acceleration.

Although the qualitative idea of flame acceleration was suggested by Shelkin already in the 1940’s [7,8], the quantitative theory of the process has been developed and supported by extensive numerical simulations only recently, by Bychkov et al. [10,11]. Among other results, the theory [10,11] predicted powerful exponential flame acceleration in micro-

(3)

channels (at least at the initial, almost isobaric stage of the process), and a decrease of the scaled acceleration rate with the channel width characterized by the Reynolds number

Re≡SLR , where S is the laminar flame speed, L R is the channel half-width (radius), and ν is the kinematic viscosity. The theoretical predictions [10,11] have been supported by later experiments on DDT in micro-channels with diameters about 1 mm and below [12,13]. At the same time, the experiments [12,13] have demonstrated some features of flame dynamics different from the theoretical predictions [10,11], such as a linear regime of flame acceleration instead of the exponential one. Reference [5] has explained the difference by the influence of gas compression effects ignored in [10,11], which moderate the acceleration process and make it linear for sufficiently large values of the flow Mach number, see also Ref. [14] for the theoretical explanation of this effect. By the end of the acceleration process, the flame propagation speed saturates to the Chapman-Jouguet deflagration speed [15-17], unless an explosion of the fuel mixture happens earlier.

Thermal losses to the walls, which are inevitably present in any experiment on DDT, is another effect expected to work against the flame acceleration mechanism. For example, in channels with isothermal cold walls, the thermal losses eventually cool the burnt gas down to the wall temperature, thus reducing effectively the effect of thermal expansion as the key element of the flame acceleration mechanism. In spite of the obvious importance of the thermal losses, quite surprisingly, not a single theoretical work on DDT has yet considered the influence of this effect on flame acceleration. At the same time, in the presence of strong thermal losses, one may question the very possibility of flame acceleration and DDT in channels with cold walls. Indeed, one should naturally expect a dominating role for thermal losses in sufficiently narrow channels, which may not only stop flame acceleration, but even lead to complete extinction of the burning process; the problem is of special importance for micro-combustion applications [18,19]. In wide channels the role of thermal losses decreases, but the flame acceleration mechanism becomes weaker too with the increase of the Reynolds number [10,11], and it remained unclear which of these two effects prevails.

In the present work we investigate the influence of thermal losses on flame acceleration in channels with cold isothermal walls. The problem is solved by using direct numerical simulations of the complete set of the Navier-Stokes combustion equations. We obtain flame extinction for narrow channels due to heat losses. However, for sufficiently wide channels, flame acceleration is found even for the conditions of cold isothermal walls. In that case flame accelerates in the linear regime, which is quite different from the powerful exponential acceleration predicted theoretically and obtained computationally for the adiabatic channels [10,11]. At the same time, the linear regime is consistent with experimental observations [12,13], which inevitably involve thermal losses to the walls. Consequently, the present work reconciles the theoretical/computational [10,11] and experimental [12,13]

endeavors, clarifying the discrepancy between them, because the boundary conditions in the practical reality are neither adiabatic, nor isothermal, but in between them. The effect of the Reynolds number on the manner of the flame acceleration is investigated.

Basic equations and the problem geometry

We consider a flame front propagating in a long 2D channel from the closed end to the open one as illustrated in Fig. 1 for the case of adiabatic walls. The problem formulation is quite similar to that of Ref. [10], with one important difference that here we study flame dynamics in channels with cold isothermal walls instead of the adiabatic ones investigated in Ref. [10], thus focusing on the influence of thermal losses to the walls on flame acceleration. The case of adiabatic walls will be also considered for the sake of comparison. We study the flame dynamics by using direct numerical simulations of the Navier-Stokes combustion equations

(4)

Figure 1. Temperature distribution (in K) for a flame propagating from the closed channel end ( =z 0) for adiabatic walls with Re =10 at the time instant SLt/ =R 0.5, see the text for other simulation parameters.

( )

=0

∂ + ∂

i i

x u

tρ ρ , (1)

( ) (

+ ,

)

, =0

∂ + ∂

j i j i j i j

i uu P

u x

t ρ ρ δ γ , (2)

2 0 1 2

1

, ⎟=

⎠

⎜ ⎞

⎝

⎛ + + −

∂ + ∂

⎟⎠

⎜ ⎞

⎝

⎛ +

j i j i j j i i

i j

i uh uu u q u

u x

t ρε ρu ρ ρ γ , (3)

( )

Y x uY xY Y

(

E R T

)

t i i i ⎟⎟=− Ra p

⎠

⎞

⎜⎜⎝

⎛

− ∂

∂ + ∂

∂ exp

Sc τ

ρ ρ ζ

ρ , (4)

where ! is the density, Y is the mass fraction of the fuel mixture, ε =QY +CVT is the internal energy, h=QY+CPT is the enthalpy, Q is the energy release in the reaction, C ,V CP are the heat capacities at constant volume and pressure. We consider a single irreversible Arrhenius type reaction of the first-order with the activation energy E and the a constant of time dimension τR. The stress tensor γi,j and the energy diffusion vector q are i

⎟⎟

⎠

⎞

⎜⎜

⎝

⎛

− ∂

∂ +∂

= ∂ ij

k k i

j j i j

i x

u x

u x u

,

, 3

2 δ

ζ

γ , (5)

⎟⎟⎠

⎞

⎜⎜⎝

⎛

∂ + ∂

− ∂

=

i i

P

i x

Y Q x T q C

Sc

ζ Pr , (6)

where ζ ≡ρν is the dynamic viscosity, Pr and Sc are the Prandtl and Schmidt numbers, respectively. The gas mixture is a perfect gas of a constant molecular weight

kg/mol 10

9 .

2 × 2

m= with CV =5Rp/2m, CP =7Rp/2m, where RP 8.31J

(

molK

)

1 is the perfect gas constant. The equation of state is

m T R

P= ρ p / . (7)

We consider a flame propagating in a two-dimensional tube of half-width R with non-slip,

=0

u , boundary conditions at the walls. Both isothermal T =Tf and adiabatic nˆ⋅ T∇ =0, (where is a normal vector to the wall ) walls are considered and compared. We take the initial pressure and temperature of the fuel mixture Pf =105Pa and Tf =300K, respectively. The thermal and chemical parameters of the fuel mixture were chosen to reproduce the most important properties of methane and propane laboratory flames. We use the dynamic viscosity ! =1.7!10"5N s/m2 and the Prandtl number Pr =1. To avoid the thermal-diffusion instability we take unit Lewis number Le ! Pr Sc =1. The activation energy

(5)

was Ea =32RpTf. We take the planar laminar flame speed SL =34.7cm/s corresponding to the initial Mach number 103. The flame thickness in our calculations is defined conventionally as

L

f S

L ≡ν/Pr ; (8)

then the Reynolds number associated with the flame speed ReSLR/ν indicates the channel half-width scaled by the flame thickness

Lf

R Pr/

Re = . (9)

Still, we would like to stress that the Reynolds number associated with the flow may be much higher due to the flame acceleration. Thermal expansion in the burning process is determined by the energy release in the reaction and characterizes the density ratio of the fuel mixture to the burnt gas Θ≡ ρf /ρb; we took Θ=8, typical for methane and propane burning.

We use a two-dimensional Eulerian code developed at Volvo Aero. The code is robust and it was utilized quite successfully in studies of laminar burning, hydrodynamic flame instabilities, development of corrugated flames, flame acceleration, and related phenomena.

The numerical scheme of the code and the computational methods were described in details in our previous papers, see, e.g., [5,6,9-11]. In the present simulations, we considered different channel half-widths 5Lf <R<25Lf, and took the tube length much bigger than the tube width, ~(100−1000)R, dynamically changing as the flame propagates. At such large values, variations of the tube length did not influence the simulation results. We use a rectangular grid with the grid walls parallel to the coordinate axes. To perform all the calculations in a reasonable time, we made the grid non-uniform along the z -axis with the zone of fine grid around the flame front. In that zone the grid size was 0.2Lf in the z -direction, which allowed us to resolve quite well the internal flame structure. Outside the region of fine grid the mesh size grows gradually with 2% change in size between the neighbouring cells; we employed the same method successfully in [10] to study flame acceleration in channels with adiabatic walls. In order to keep the flame in the zone of fine grid, we applied an adaptive mesh moving together with the flame. Along the x -axis we used a uniform grid. The number of cells in the z -direction was different for different tube widths: the wider the tube, the larger the number of cells. To keep the simulation time reasonable in wide tubes we took up to 110 cells in x - direction, so that the grid size was comparable to L . By using such a grid we were able to f resolve quite well the zone of large velocity gradients close to the walls. To check if the number of cells was sufficient for the problem, we have performed test simulation runs with number of cells increased 3 times in x-direction and obtained the same results within an accuracy of about 5%. This value may be considered as the numerical accuracy of our computations. Similar to [10], we used the Zeldovich-Frank-Kamenetsky solution for a planar flame front as an initial condition. The planar flame front was created at a distance 6Lf from the closed tube end. We kept non-reflecting boundary conditions at the open end of the tube as described in [10]. Using such conditions we avoided reflections of weak shocks and sound waves from the open end, which otherwise might influence burning and the process of flame acceleration.

Simulation results

For the fixed density ratio Θ=8 used in the present work, the main simulation parameter is the Reynolds number, which indicates the scaled half-width of the channel, Eq. (9).

Depending on the Reynolds number, we obtain qualitatively different flame behavior:

extinction for narrow channels or acceleration for sufficiently wide channels. In particular, Figure 2 shows evolution of the temperature pattern in the simulation run for Re =5

(6)

Figure 2. Temperature distribution (in K) changing in time for the case of isothermal walls with Re =5 at the time instantsSLt/ =R 0;0.24;0.48.

corresponding to the tube half-width R 5= Lf at the time instants SLt/ =R 0;0.24;0.48. As we can see, in that case the flame is not ignited and the initially heated region by the tube end gets eventually cooled down by the walls. This result agrees well with the commonly accepted knowledge that a propagating flame stands away from a cold wall with the quenching region about six times the laminar flame thickness [20,21]. Taking a wider tube

Lf

R 10= with Re =10, we manage in igniting a self-sustained flame, which starts propagating for a while as shown in Fig. 3. Still, in contrast to the adiabatic case illustrated in Fig. 1 we observe relaxation of the burnt gas temperature down to the wall temperature because of the heat losses. The relaxation process is especially strong by the tube end, although we observe similar effects by the side walls too with the quenching distance between flame “skirt” and the walls about ~4Lf, in reasonable agreement with [20,21]. In the case of Re =10, the losses to the cold walls eventually prevail over the flame propagation and, after an initial short acceleration, we find a slowing down of the flame front, as shown in Fig. 4, with subsequent flame quenching. For comparison, Figure 4 presents also the propagation speed of the flame tip for the case of adiabatic walls with the same Reynolds number,

-1 0 1

x / R

30 25

20 15

10 5

0

z / R

20001500 1000500

Figure 3. Temperature distribution (in K) for a flame in a channel with isothermal walls for Re =10 at the time instant SLt/ =R 1.6.

(7)

Figure 4. Propagation speed of the flame tip versus the scaled time SLt/ for R Re =10for isothermal and adiabatic walls.

Re =10, which exhibits powerful acceleration in agreement with the previous works [5,10].

By increasing the tube width further, with Re >10, we finally obtain flame acceleration in channels with cold isothermal walls too; the snapshots of the characteristic temperature patterns are presented in Fig. 5 for Re =15 at the time instants SLt/ =R 0;0.2;0.4;0.6;0.8. Similar to Fig. 3, in Fig. 5 we also observe temperature relaxation by the tube end and the side walls; however this time the heat loss is not strong enough to stop the self-sustained flame acceleration - the curved flame front produces more gas volume per unit time than the losses can “remove” by relaxation of the burnt gas temperature. The propagation speed of the flame tip for Re =15 and isothermal channel walls is shown in Fig. 6 versus the scaled time

R t

SL / ; the associated adiabatic case is also presented on the figure for comparison.

According to Fig. 6, after some initial transitional time, the flame front accelerates in approximately linear regime in the channel with isothermal walls, which may be described roughly as

const R

t S a S

Utip / L = ( L / )+ , (10)

where a is the scaled flame acceleration. This linear acceleration regime differs considerably from the powerful exponential acceleration Utip /SL ∝exp(σSLt/R) predicted theoretically and validated numerically in Ref. [5] for channels with adiabatic walls. At the same time, we point out that the experiments on flame acceleration in ethylene-oxygen fuel mixtures in micro-channels have reported a linear acceleration law rather than the exponential one [12,13]. Although efficiency of heat losses to the walls has not been directly investigated in the experiments [12,13], still, in the realistic combustion process in channels some losses are inevitable. It would be natural to assume that the experimental situation of Ref. [12,13]

corresponds to something intermediate in-between the asymptotic cases of isothermal cold walls and adiabatic ones. The linear regime has been also observed at the developed stages of flame acceleration in adiabatic channels when the gas compressibility effects become important and the Mach number of the flow approaches unity [5]. Besides, linear flame acceleration has been encountered in numerical modelling even at the initial stages of DDT in

(8)

Figure 5. Temperature distribution (in K) for the flame acceleration in the channel with isothermal walls for Re =15 at the time instants SLt/ =R 0;0.2;0.4;0.6;0.8.

Figure 6. Propagation speed of the flame tip versus the scaled time SLt/ for R Re =15for isothermal and adiabatic walls.

(9)

Figure 7. Propagation speed of the flame tip versus the scaled time SLt/ for R Re = 15;17.5; 20for isothermal walls.

adiabatic channels provided for a sufficiently small density ratio or sufficiently large Reynolds number [22]. By comparing all these cases, one can presumably discuss a common tendency of the linear flame acceleration regime replacing the adiabatic one as soon as the acceleration mechanism gets moderated and weakens by some reason. One is also tempted to treat the linear regime of flame acceleration, Eq. (10), as the first term in the Taylor expansion for the weak exponential acceleration Utip/SL∝exp(σSLt/R)≈1+σSLt/R with a small acceleration rate σ <<1. Still, one has to be cautious with such a treatment as the linear acceleration regime has been observed up to the instants with noticeable increase of the flame propagation speed, when the term σSLt/R is not small any more.

By investigating the flame acceleration process in channels of different width, we have obtained that acceleration goes slower for wider tubes as shown in Fig. 7. The scaled acceleration a extracted from the plots of Fig. 7 according to Eq. (10) is presented in Fig. 8 versus the Reynolds number; the decrease of the scaled flame acceleration with Re may be observed for Re ≥15 (filled markers). This tendency agrees qualitatively with the results obtained for the channels with adiabatic walls [10]; we remind that the theory of Ref. [10]

predicted asymptotic decrease of the acceleration rate σ with the Reynolds number as Re

2 / Θ

σ = in wide tubes with Re>> 4Θ. The scaled flame acceleration at the developed stages of the process decreases with the Reynolds number too, as found in Ref. [5]. Thus, all these cases demonstrate the same qualitative tendency of the flame acceleration mechanism becoming weaker for wider tubes with smooth walls.

The domain of narrow tubes Re <10 in Fig. 8 corresponds to flame quenching and extinction. The border case of Re =10 is indicated by the empty marker with some minor acceleration extracted for the initial stage of the flame propagation. So, unlike the adiabatic case, the flame acceleration in channels with cold isothermal walls depends on the Reynolds number in some non-monotonic way with the maximal acceleration attained for 10<Re<15.

(10)

Figure 8. Scaled flame acceleration, Eq. (11) versus the Reynolds number for channels with isothermal walls.

Summary

In the present work we have investigated the problem of flame dynamics – extinction versus acceleration – in narrow long channels with cold isothermal walls with flame propagating from the closed channel end to the open one. The problem is very important in the context of the DDT studies; in particular, it is needed to reduce the gaps between the theoretical models and simulations for the flame acceleration in channels with adiabatic walls presented so far [5,10,11] and the experimental results on DDT in micro-channels [12,13]. The problem is solved by means of direct numerical simulations of the complete set of equations of combustion and compressible hydrodynamics, including transport properties (heat conduction, diffusion and viscosity) and the Arrhenius chemical kinetics.. We have obtained qualitatively different flame behavior in sufficiently narrow and wide channels. For the channel half-width smaller than about ten flame thicknesses heat losses to the cold walls prevail and burning is eventually quenched. In wider channels flame acceleration is found even for the conditions of cold walls in spite of the heat loss. In that case flame accelerates in the linear regime, which differs considerably from the powerful exponential acceleration predicted theoretically and obtained computationally for the adiabatic channels in Ref. [10].

At the same time the linear regime is consistent with the previous experimental observations [12,13], which inevitably involve thermal losses to the walls. We also find that the scaled flame acceleration decreases with the Reynolds number of the flow for channels with cold isothermal walls; this tendency is qualitatively similar to the case of adiabatic channel walls studied previously.

References

[1] Roy G.D. , Frolov S.M. , Borisov A.A., Netzer D.W., “Pulse detonation propulsion:

challenges, current status and future perspectives”, Prog. Energy Combust. Sci. 30: 545- 672 (2004)

[2] Ciccarelli G., Dorofeev S., “Flame acceleration and transition to detonation in ducts”, Prog. Energy Combust. Sci. 34: 499-550 (2008)

(11)

[3] Dorofeev S., “Flame acceleration and explosion safety applications”, Proc. Combust.

Inst. 33: 2161-2175 (2011)

[4] Kuznetsov M., Alekseev V., Matsukov I., Dorofeev S., ”DDT in smooth tube filled with hydrogen-oxygen mixture”, Shock Waves 14: 205-215 (2005)

[5] Valiev D., Bychkov V., Akkerman V., Eriksson L.-E., “Different stages of flame acceleration from slow burning to Chapman-Jouguet deflagration”, Phys. Rev. E 80:

036317 (2009)

[6] Valiev D., Bychkov V., Akkerman V., C.K. Law, Eriksson L.-E., “Flame acceleration in channels with obstacles in the deflagration-to-detonation transition” Combust. Flame 157: 1012-1021 (2010)

[7] Shelkin K.I., ”Influence of tube roughness on the formation and detonation propagation in gas”, Zh. Eksp. Teor. Fiz. 10: 823-828 (1940)

[8] Zeldovich Ya. B., Barenblatt G.I., Librovich V.B., Makhviladze G.M., Mathematical Theory of Combustion and Explosion, Consultants Bureau, 1985

[9] Bychkov V., Valiev D., Eriksson L.E., ”Physical mechanism of ultrafast flame acceleration”, Phys. Rev. Lett. 101: 164501 (2008)

[10] Bychkov V., Petchenko A., Akkerman V., Eriksson L.E., “Theory and modelling of accelerating flames in tubes” Phys. Rev. E 72: 046307 (2005)

[11] Akkerman V., Bychkov V., Petchenko A., Eriksson L.E., ”Accelerating flames in cylindrical tubes with nonslip at the walls”, Combust. Flame 145: 206-219 (2006)

[12] Wu M., Burke M., Son S., Yetter R., “Flame acceleration and the transition to detonation of stoichiometric ethylene/oxygen in microscale tubes”, Proc. Combust. Inst.

31: 2429-2436 (2007)

[13] Wu M., Wang C.Y., “Reaction propagation modes in millimeter-scale tubes for ethylene/oxygen mixtures”, Proc. Combust. Inst. 33: 2287-2294 (2011)

[14] Bychkov V., Akkerman V., Valiev D., Law C.K., ”Role of compressibility in moderating flame acceleration in tubes”, Phys. Rev. E 81: 026309 (2010)

[15] Chu R., Clarke J., Lee J.H., “Chapman-Jouguet deflagrations”, Proc. Roy. Soc. London A 441: 607-623 (1993)

[16] BychkovV., Valiev D., Akkerman V., Law C.K., “Gas compression moderates flame acceleration in deflagration to detonation transition”, Combust. Sci. Techn. 184: 1066- 1079 (2012)

[17] Valiev D., Bychkov V., Akkerman V., Eriksson L.E., Law C.K., ”Quasi-steady stages in the process of premixed flame acceleration in narrow channels”, Phys. Fluids 25:

096101 (2013)

[18] Ju Y., Maruta K., “Microscale combustion: technology development and fundamental research”, Prog. En. Combust. Sci. 37: 669-715 (2011)

[19] Kaisare N., Vlachos D., “A review on microcombustion: fundamentals, devices and applications”, Prog. En. Combust. Sci. 38: 321-359 (2012)

[20] Ferguson, C. R., and Keck, J. C., “On laminar flame quenching and its application to spark-ignition engines” Combust. Flame 28:197-205 (1977)

[21] Daou J., Matalon M., “Influence of conductive heat losses on the propagation of premixed flames in channels”, Combust. Flame 128: 321-339 (2002)

[22] Demirgok B., Ugarte O., Valiev D., Akkerman V., “Effect of thermal expansion on flame propagation in channels with nonslip walls”, Proc. Combust. Inst. (2014), in press.

References

Related documents

The partial differential equations are discretised using a finite difference or finite volume method on a structured grid, which is a common approach in the CFD community.. In

Study III investigated accompanying symptoms, such as cold extremities, constipation, pain, sleeping disorders, and well-being, and their treatment to determine (i) whether

A larger proportion of non-walking children than of walking children was reported to have accompanying neurological impairments, cold extremities, pain,

The results of this study do not support higher prevalence of psychiatric disorders among MND patients, in fact the occurrence is slightly lower than that for the general

T ABELL 6.5: A NTAL AKUTA ORDER SOM LAGTS TILL CENTRALLAGRET FRÅN VARJE TERMINAL ENLIGT UHS

Even though this kind of coercion undermines the will of the demented person, some argue that it is justified when it takes place for the individual's own good, since he or she

The main results produced by the efforts in this thesis are: x the identification of activities needed for IT security assessment when using IT security metrics, x a method

Vår andra frågeställning var om det finns en koppling mellan studenter som är frekvent tillgängliga via mobiltelefonen och deras generella hälsa.Vår hypotes var att