You are on page 1of 9

Nuclear Engineering and Design 263 (2013) 500508

Contents lists available at SciVerse ScienceDirect

Nuclear Engineering and Design


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

Computational and experimental prediction of dust production in pebble bed reactorsPart I


Maziar Rostamian a, , Gannon Johnson a , Mie Hiruta a , Gabriel P. Potirniche a , Abderra M. Ougouag b , Joshua J. Cogliati b , Akira Tokuhiro a
a b

Department of Mechanical Engineering, University of Idaho, 1776 Science Center Drive, Idaho Falls, ID 83401, USA Idaho National Laboratory, 2525 N Fremont Avenue, Idaho Falls, ID 83401, USA

h i g h l i g h t s

A nonlinear dimensionless wear coefcient is theoretically proposed. A material constant for the relation of asperity height and wear is introduced. A nonlinear modication of Archard wear formula is proposed. The graphite wear dust production in a typical pebble bed reactor is predicted. Experimental and computational wear results for graphite are presented.

a r t i c l e

i n f o

a b s t r a c t
This paper describes the computational modeling and simulation, and experimental testing of graphite moderators in frictional contacts as anticipated in a pebble bed reactor. The potential of carbonaceous particulate generation due to frictional contact at the surface of pebbles and the ensuing entrainment and transport into the gas coolant are safety concerns at elevated temperatures under accident scenarios such as air ingress in the high temperature gas-cooled reactor. The safety concerns are due to the documented ability of carbonaceous particulates to adsorb ssion products and transport them in the primary circuit of the pebble bed reactor, thus potentially giving rise to a relevant source term under accident scenarios. Here, a nite element approach is implemented to develop a nonlinear wear model in air environment. In this model, material wear coefcient is related to the changes in asperity height during wear. The present work reports a comparison between the nite element simulations and the experimental results obtained using a custom-designed tribometer. The experimental and computational results are used to estimate the quantity of nuclear grade graphite dust produced from a typical anticipated conguration. In Part II, results from a helium environment at higher temperatures and pressures are experimentally studied. 2013 Elsevier B.V. All rights reserved.

Article history: Received 3 November 2012 Received in revised form 16 April 2013 Accepted 19 April 2013

1. Introduction The vast majority of todays nuclear reactor designs, predominantly light water reactors, have xed fuel and coolant congurations. The pebble bed reactor (PBR), having neither of these, uses nuclear fuel encased in a random conguration of spherical fuel elements called pebbles in a cylindrical bed. While this design is capable of online refueling and high output temperatures, its design also creates specic modeling and design challenges for safety and licensing (IAEA, 2003). Of these challenges, one that is

Corresponding author. Tel.: +1 208 310 9555. E-mail addresses: mrostamian@asme.org, mzram 22@yahoo.com (M. Rostamian). 0029-5493/$ see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.nucengdes.2013.04.019

specic to Graphite-Moderated Gas Cooled Reactors (GM-GCR) is the phenomenon of carbonaceous particulate generation due to frictional contact of graphitic components. Due to the number of moving carbonaceous components (i.e. fuel elements or pebbles), this is primarily a concern for pebble bed reactors for safety and operation. Safety concerns are focused around understanding transport and deposition of ssion and activated products, and how the dust will be transported during a Depressurized Loss-Of-Flow Cooldown accident (DLOFC). In particular, under DLOFC and similar hypothetical accidents, the main issues are as follows: quantication of the inventory of dust, dust re-suspension off of surfaces in the primary circuit, dust departure from a breached primary circuit, contamination levels associated with such an event (source term), and assessment of the additional risks posed by any energetic

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

501

mechanisms of the graphite dust suspending into the surrounding air (NUREG/CR-6944). While there is limited data on these phenomena in the published literature from past GM-GCRs and R&D, it is not sufcient for the level of details expected under modern design review and licensing requirements of a pebble bed by the United States Nuclear Regulatory Commission (US-NRC). This is further complicated by the fact that some materials in the past GM-GCRs are either not used in current designs (e.g. Inconel 617 being replaced by modern alloys), or available (e.g. nuclear graphite H-451) (Kissane, 2009). A panel of experts ranked the phenomena as medium in the Phenomena Identication Ranking Table (PIRT) and the knowledge level as low (NUREG/CR-6944). The qualitative ratings in a PIRT of high, medium, and low are used to identify the importance of expected phenomena and the level of understanding of those phenomena. The US Nuclear Regulatory Commission Regulation (NUREG) also identies a need for graphite dust production and tribological data in a helium environment, as an anticipated function of temperature, pressure, and uence. In an effort to quantify the frictional wear of materials (fuel, non-fuel) used in the pebble bed, a custom tribometer was developed to measure the wear of materials in helium atmosphere at pressures and temperatures expected in a modern GM-GCR. This paper describes the design, deployment of the device and the initial experimental data from several grades of nuclear graphite. Also, in this paper the analytical linear Archard model is modied to account for the wear of brittle materials. In an attempt, the wear coefcient in this equation, which is conventionally assumed constant, is derived as a function of the surface asperity height. The nite element computational model of contacting fuel pebbles is developed in ABAQUS.

asperities. This is also seen in the experimental results where after 60 m of frictional contact, only a small worn mass of approximately 1 g was accumulated (Johnson, 2012). Therefore, an analytical model has been established to consider the effect of asperities on graphite wear. Most brittle materials including graphite have a linear wear behavior for small sliding distances (Blau, 1992). However, for most real-life applications, the materials that are subject to frictional contact are used over long time periods, which means they undergo long sliding distances. In a previous work, it was attempted to model a 2-m sliding wear (Rostamian et al., 2012). The results were compared with Cogliatis results (Cogliati and Ougouag, 2010) and his reviews on Xiaoweis results (Cogliati et al., 2011). However, it is now attempted to model a much longer sliding length, so that the nonlinear wear map (i.e. wear versus distance variation) generated from the experimental results can be predicted. In order to capture this phenomenon, the linear Archard wear model has been examined, modied and adapted.

2.1. The linear Archard wear model The Archard wear equation is a linear model derived from pin-on-disk (perpendicular contact) and pin-on-pin (tangential contact) experiments by the French scientist, J.F. Archard, in the 1950s (Archard and Hirst, 1956). In this model, the wear volume is related to the material properties and the sliding distance in the following form FL H

V =K

(1)

2. The computational analysis Wear in metals is commonly in the form of microcracks initiating under high local plastic strains. In the contact region, the highest plastic strain is seen on the perimeter of the contact area. This is where microcracks are likely to initiate. Existence of a continuous contact stress can easily cause to propagate the microcrack. For plane stress, there are two major crack modes: mode I where a tensile stress is normal to the plane of the crack, and mode II where a shear stress acts parallel to the plane of the crack and perpendicular to the crack front. Under the frictional contact condition, mode I is the possible crack type. Microcracks that are formed can propagate only if their stress intensity surpasses the critical crack stress intensity as formulated by Irwin (1957). If this criterion is satised, the cracks usually start to propagate perpendicular to the maximum principal stress (Pook, 2007). If the contact stress continues to exist or if the segment under study goes through a high number of cycles, cracks on both sides of a 2D contact zone can meet one another at a certain depth underneath the contact zone. This mechanism produces chips, which leads to abrasive wear. Abrasion can be a form of wear in brittle materials. However, at very small inter-pebble loadings present in PBRs, this needs to be investigated. Following a previous macro-level analysis (Rostamian et al., 2012), and using eXtended Finite Element Method (XFEM), crack domains were considered on the surface of two graphite pebbles in contact. Considering the domain of applicability for the inter-pebble forces, which is a range of 1050 N (Cogliati and Ougouag, 2008; Rostamian et al., 2012), crack initiations studies were performed. However, it is revealed that the wear mechanism by plastic deformation and crack initiation has an onset of 1000 N for IG-11 graphite pebbles with a diameter of 3 cm. Therefore, it is assumed that the wear in brittle materials such as graphite under smaller loadings is caused by powder formation from the frictional contact between surface

where K is the dimensionless wear coefcient or the proportionality constant, which is a material parameter; F is the force normal to the contact point; L the sliding distance; and H the material hardness. It is seen that the wear volume is directly related to the normal forces and the sliding distance. By performing nite element simulations at a macro-level, the Hertzian normal forces can be easily determined (Johnson, 1987). As a result, the relationship between the wear volume/mass and the sliding distance can be obtained (Rostamian et al., 2010, 2011, 2012). However, this is only true for small sliding distances in the case of brittle materials. In order to account for the nonlinear wear rates or behavior of brittle materials, one needs to consider the material surface at a micro-level, where asperities are considered as the real contact points. In the following section, a micro-level study of the contact surface will be considered and the role of asperity height in presented in the newly introduced nonlinear version of the Archard model will be explained.

2.2. Micro-level considerations The height of asperities and the surface topology determine the roughness of the material surface. Let us consider two segments made of one material; one of a rough surface and one of a smooth surface. It is obvious that the material of a rougher surface has a higher wear volume after a certain distance of sliding wear. This is also seen in the air test experiment results from the tribometer at room temperature. The wear ring around the pebbles is seen to have been smoothed after the pebbles have been in contact (see Section 3 and Fig. 6). Therefore, by a closer examination of the Archard wear model, we strive to assess the dependence of wear volume on the asperity height.

502

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

2.3. A new wear correlation Based on the microlevel consideration above, we consider the = dV/dt ) form of the Archard equation: volumetric wear rate (V FL =K V H (2)

where = (hi H )/(Lmax P ). Now, by introducing the following asperity height power-law function, we strive to model the effect of surface smoothing: (h) = R(h)
n

(10)

Introducing the asperity height into this equation, we approximate the volume by V = hA. Therefore, =K hA FL H (3)

where R is an asperity model constant to be obtained by comparing the analytical model with the computational data and n is the exponent of the power-law function, also to be determined through a t of the computational results. Substituting the asperity height function (h) into Eq. (9), we will have:

where A is the contact area. The negative sign denotes that the time rate change of asperity height is negative, representing a decrease in asperity height with time. It should be noted that the contact area can vary based on the component shape and the contact congurations. In order to generalize this equation for any contact area and component shape, we consider the contact stresses instead of normal forces. Then, Eq. (3) takes the form: =K h PL H (4)

dh = K, dL

h = hi

(11a)

dh n = K (h ) = KR (h ) , h < hi (11b) dL
K

where K = KR(h ) is a varying dimensionless wear coefcient, which depends on the variation of asperity height with wear length and on the material properties. By rearranging and integrating, the varying wear coefcient, K , can be derived as a function of the sliding length, L*, and the constants R and n. There follows, K = KR [(1 n)L ]n
1/(1n)

where P is Hertzian contact stress. By performing computational simulations in ABAQUS, contact stresses can be easily obtained at is sliding distance rate or relative each node in the mesh. In Eq. (4), L velocity at contact. However, the time rate of change in the asperity , is unknown. To eliminate this unknown, we consider the height, h following: h = h(t ) = h(L(t )) (5)

(12)

which denotes that the dependence of asperity height on time can be formulated as the dependence of asperity height on wear length. Here, the gas adsorption effects on graphite surface are not consid , the partial derivative of h with respect to ered. In order to nd h time is used: dh dh dh dL h = dh L = = h dt dL dt dL dL L Therefore, H dh =K P dL (7) (6)

This is still a linear equation for the rate of change of the asperity height. However, it can now be argued that the asperity height which decreases with the sliding length can start to inuence the material constant K. This material property is commonly considered as a constant when the Archard equation is implemented for wear calculations. In this work, we introduce a function, which take into account the effect of asperity height variations. This function incorporates the effect of asperity height decrease into the linear Archard wear model for higher sliding distances. When the asperity height decreases from its initial height, the surface becomes smoother, which indicates that the wear rate should decrease. Before introducing the asperity height function, we consider the following normalized parameters to simplify the equations. L = L , Lmax h = h hi (8)

The wear coefcient thus derived depends on two parameters: the asperity model constant, R, and the power law coefcient n. These material constants are determined through a tting process described here. The nite elements computational model presented in Section 2.4 allows the determination of asperity height as a function of wear length. The FEA model is used repeatedly to obtain a function describing the height, h, of a given asperity versus the wear length, L, as it is varied between 0.0 and 2.0 m. The derivative of that function is proportional to K . Knowing the computed (L, K ) pairs and introducing the L values into Eq. (12) for K , model values for K are obtained for various values of the parameters R and n. This process allows the formulation of a tting problem for the R and n parameters, based on minimizing the difference between the computed K and the model K . The values of R and n that are retained in this work are the rst pair found that results in a maximum absolute error below 106 for K . The parameters obtained in this way from computational model data up to 2.0 m of wear length are then used for predicting the wear performance for up to 1200 m of wear length. The analytical model predicts experimental data with remarkable accuracy and delity as detailed later through comparison with the experimental results. All the tting work described here assumed the properties of graphite IG-11. 2.4. The computational analysis The nite element software ABAQUS v6.11 has been used to conduct simulations of the micromechanics of wear (ABAQUS users manual). Primarily, 3D spherical asperities were considered to be modeled using the implicit module of ABAQUS known as the ABAQUS Standard. However, given the demanding nature of contact conditions, the nonlinear geometry of the asperities, severe deformations, and most importantly, the overlapping of asperities, the implicit module of ABAQUS faced convergence issues. An alternative was taken by applying the same methodology in the explicit solver of ABAQUS known as the ABAQUS Explicit, which is very robust in handling contact problems. This module however demands higher run times. As a result, it was decided to switch from spherical to cylindrical asperities.

where Lmax is the maximum sliding length and hi is the average asperity height of the component unworn surface. Therefore, Eq. (7) takes the form dh =K dL (9)

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

503

occurrence of hourglass deformations in an analysis can invalidate the results and should always be minimized. Therefore, hourglass control is used to overcome this deciency. This formulation provides improved coarse mesh accuracy with slightly higher computational cost and performs better at high strain levels. Hourglass control is an embedded feature in ABAQUS v6.11. Plane strain is simulated in this 3D model by constraining the front and back planar faces of the asperity as symmetry (i.e. uz = 0, urx = ury = 0). In this way, the model behaves as a plane strain slice. A prescribed velocity was primarily used for the movement of the upper asperities, and the lower asperity was kept xed as its base. Hence, to simulate the movement of both pebble surfaces, the relative tangential velocity is applied to the upper asperity. In a later attempt, it was found out that prescribing velocities to both upper and lower asperities in opposite directions provides more realistic results for the two asperities. Therefore, the actual tangential velocity of pebbles at contact was applied to upper and lower asperity bases. 2.4.2. Mass scaling Modeling using an explicit procedure can take a considerable amount of time for the simulation to converge. This is due to the fact that the minimum stable time increment is small. Using appropriate mass scaling, the computational efciency is improved while retaining the necessary degree of accuracy required for a particular class of problems. Mass scaling was introduced into the material model in multiple steps. This means that in every step, an articial mass was added and the simulations were performed. After assessing the results, if inertial effects were not seen as compared with the results of no articial mass, then some more mass was introduced. Otherwise, mass scaling would not be continued. 2.4.3. Material model The material properties for nuclear grade graphite IG-11 were chosen for the asperity model. Here we have used a damage model and have included damage evolution to visualize the degradation of the fully damaged elements. Material failure refers to the complete loss of load-carrying capacity, which results from

Fig. 1. A meshed asperity of radius 3 m and thickness 0.04 m.

2.4.1. Finite element model Fig. 1 shows the nite element mesh used for the cylindrical asperity model. This asperity model was made of linear hexahedral elements (C3D8R). R stands for reduced-integration, which uses fewer number of Gaussian points when solving the integral. The advantage of the reduced integration elements is that the strains and stresses are calculated at the locations that provide optimal accuracy; the so-called Barlow points (Barlow, 1976). A second advantage is that the reduced number of integration points decreases CPU time and storage requirements. In the case of cylindrical asperities which is considered in this study, it would be common sense to conduct the simulation in 2D considering shell elements. However, since element removal is sought in these simulations, the elements which are originally within the interior of the asperity can start to interact with the opposing asperity as a contact surface. This feature is only avalibale in general contact algorithm in ABAQUS, which requires the use of 3D solid elements. Thus, a 3D model of cylindrical asperities was constructed. Despite being robust for large deformations and saving extensive amounts of run time, the reduced-integration solid and shell elements used in ABAQUS are prone to zero-energy modes. These modes, commonly known as hourglass modes, are oscillatory in nature and result in mathematical states that are physically unfeasible. They typically have no stiffness and give a zigzag appearance to the mesh known as hourglass deformations as seen in Fig. 2. The

Fig. 2. A schematic of hourglass deformation.

Fig. 3. Elasticplastic behavior with hardening for nuclear graphite IG-11.

504

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

progressive degradation of the material stiffness. An undamaged constitutive behavior of the material (i.e. elastic-plastic with hardening) is determined (Yokoyama et al., 2008) as seen in Fig. 3. Then, a damage initiation criterion is determined (point A). This is followed by damage evolution (paths AB) and nally the choice of element removal (point B). In the absence of any experimental data for the damage softening part of the curve (AB), a simple linear damage evolution law was used to degrade the material stiffness to zero where the equivalent total strain reaches 0.05 (point B in Fig. 3). When an element reaches point B on the curve, its stiffness has fully degraded, and the element is removed from the mesh. 2.4.4. The nite element analysis The nite element analysis of asperity contact in ABAQUS Explicit revealed that asperities undergo very high deformations. As the material reaches a total strain of 0.05, the elements are removed and the degradation leads to material removal from the asperity. In Fig. 4, the degradation and material removal from asperities of 3 m high are presented. The height of the worn asperity from the rst simulation was used to perform the next simulation on this asperity after one full rotation of the graphite pebble. This process continued to derive the right hand side of Eq. (11b), which is needed to derive the asperity model constant, R. 3. The experiment The basic sample design conceived for testing to mimic the geometry of the pebble bed with the apparatus is a graphite disk machined into 2.54 cm tall disks with a 3 cm outer contour (Johnson, 2012). 3.1. The experimental apparatus The test data used in this paper is at room temperature and atmospheric pressure. This is because the material properties for nuclear graphite are only available at room temperature. Also, there is no test data for graphite matrix or graphite at the temperatures and pressures expected in a pebble bed. As a result, a custom test apparatus was designed. In the following section, the experimental setup is explained in brief. The motion requirements for rolling and sliding between two spherical objects led to the selection of a twin disk tribometer. Because no off-the-shelf twin-disk tribometer could operate at the temperature and pressure requirements of PBR, a custom apparatus had to be designed. However, only the results at room temperature and atmospheric pressure are reported in the present paper. High temperature and pressure results are presented in a parallel but different work (Part II). The samples were also measured before and after the experiments on an analytical balance for a separate method to measure the average wear rate of the samples for comparison. A schematic of the tribometer with the pebbles in contact is presented in Fig. 5. The Scanning Electron Microscopy (SEM) images (Fig. 6) from the graphite samples, after an approximate sliding distance of 1000 m show a wear scar of 1.7 mm wide with a smooth and clean surface. This is while outside of wear scar, a rough surface from the fabrication process is observed. This indicates that the asperities were polished and leveled by the rotating motion.

Fig. 4. Equivalent plastic strain contour: the degradation and material removal process from asperities at: (a) t = 0.000325 s, (b) t = 0.0013 s, (c) t = 0.002275 s and (d) t = 0.00325 s for a tangential speed of 0.08 m/s.

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

505

Fig. 5. Tribometer (left) and its schematic (right) with pebbles in contact and the modular analytical balance.

4. Results and discussion The wear mass results from the tribometer are presented in Fig. 7. The wear mass was recorded every 200 m. It is clearly seen that the primary wear rates are reduced as the sliding length increases. The wear coefcient obtained from the analytical model is compared in Fig. 8a to the computational results, and in Fig. 8b to the experimental data. The analytical model exponent and the asperity model constant were derived by tting with the computational results. The model thus obtained is then compared with the experimental data for verication purposes. The exponent, n, is computed to be 1.38 and the dimensionless microlevel asperity model material constant, R, is 2.5 104 . It is clearly observed in Fig. 8b that the proposed analytical model, which stems from the computational data, predicts a higher overall wear coefcient than the actual wear coefcient derived from experimental data. This is because the natural lubricating properties of graphite surface have not been considered (Czichos, 1978). As explained earlier, when two graphite surfaces are in contact, powder formation is the main form of wear. The powder is then capable of lling in the surface valleys, which causes the surface

Fig. 6. The SEM images of graphite sample (a) wear scar zone (b) interface between the scar (lower) and rough zone (upper).

Fig. 7. Wear mass loss from the tribometer air-tests for 1100 m.

506

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

Fig. 8. Comparison of analytical K with (a) computational results and (b) experimental data with respect to the dimensionless sliding length, L*. Fig. 9. Instantaneous and cumulative wear mass for one graphite pebble with respect to wear length, L.

to become smoother than what is incorporated into our ABAQUS model. A plausible explanation is that the uneven distribution of asperity height on an actual surface could cause a different nonlinear wear coefcient than what is predicted here. The minimum (nal) asperity height, as predicted by this model, is 12.45% of the initial average asperity height. This indeed indicates that the surface has become smoother and thus the wear coefcient smaller. Using the new varying wear coefcient, the wear results, which were predicted before based on a constant coefcient, can now be rectied. As seen in Fig. 9a, the initial wear rate starts to decrease by wear distance and the cumulative wear is extremely small for air-tests. The predicted wear mass using the proposed analytical model is compared to the experimental results in Fig. 9b. As anticipated, due to the difference between the predicted and actual dimensionless wear coefcients, the predicted wear mass is approximately 20% higher than the actual wear mass. It is clearly observed in Fig. 9b that the proposed wear model shows a signicant improvement in comparison to the Archard wear equation. The error bars indicated on the graph in Fig. 9b are from the uncertainty associated with the scale with which the experimental data was recorded. This indicates that the analytical model has predicted reliable results that t in within the uncertainty limits.

The varying wear rate of graphite is shown in Fig. 10. It is seen that the wear rate decreases as graphite asperity height decreases with increasing wear length. The reported wear rate in (Johnson, 2012) was a constant 8.45 108 . This value is the same as the initial rate predicted in the present study. However, in the present study, the rate is shown to be varying over time. It must be noted that the present results are for air test at room temperature and atmospheric pressure. Using the dimensionless wear coefcient K obtained in this work, it is attempted to derive the wear mass for one pebble through 2.06 m of HTR-10 wear length as in (Rostamian et al., 2012). By performing wear calculations as carried out by Cogliati et al. (2011) based on the air-test results of Xiaowei et al. (2005a), the wear mass produced by inter-pebble forces (Cogliati and Ougouag, 2008) is calculated. Estimates for HTR-10 and AVR are reported using the present model. The wear mass for one pebble is estimated to be 0.235 mg. For HTR-10 with an inventory of 27,000 pebbles and a pebble ow rate of 125 pebble per day, the wear mass is predicted to be 5.36 g/year. As explained by Rostamian et al. (2012), the estimated wear mass for AVR would then be 63.55 g/year.

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508

507

obtained from the analytical model and the experimental data. This is attributed to the lack of a surface-level lubrication component in the model, thought to be induced by gas adsorption as well as relocation of surface-adhering graphite particulates. It is known that graphite acts as a good lubricant. Thus it could be surmised that when particulates are formed due to wear, they primarily relocate in the valleys between asperities. This increases the smoothness of the surfaces, which in turn decrease the wear coefcient. The lamellar structure of graphite facilitates the establishment of smooth surfaces between sliding contacts, and the presence of adsorbed gases may further improve the quality of these surfaces. Since these artifacts are not considered in the nite element simulations, the predicted wear coefcient is slightly larger than that attributed to the experimental data. However, it is seen that the proposed model is capable of representing the nonlinear wear rates as seen in experimental results and that it falls within the limits of the experimental data uncertainty. The proposed model and the experimental results show that the wear mass for air environment at room temperature and atmospheric pressure is small and on the order of grams per year for a reactor. Therefore, it is predicted that the presence of helium environment at higher temperature and pressures, actually contributes to increased wear. It was shown by Xiaowei et al. (2005b) that the wear rate of graphite at elevated temperatures is much higher than that at room temperature. High temperature/pressure tests in helium gas were also tested using the custom-designed tribometer. The results of these studies are reported in Part II as a continuation of the work presented here.

Acknowledgement We express our gratitude to DOE, under NEUP09-151, The Experimental Study and Computational Simulations of Key Pebble Bed Thermomechanics Issues for Design and Safety for providing nancial support for this study.

Fig. 10. Wear rate for one graphite pebble with respect to wear length, L.

In Part II of this paper, helium-test experimental data at higher temperatures and pressures will be reported. It is known that higher temperature and pressure have signicant impacts on the wear properties of graphite. These effects will be discussed in Part II. 5. Conclusions In the present work, we investigated the generation of carbonaceous particulate due to frictional contact at the surface of graphite pebbles. Such particulates can give rise to safety issues at elevated temperatures, under accident scenarios for a high temperature gas-cooled reactor. Specically, the adsorbed ssion products onto the graphite dust represent a source-term concern. Here, a micro-level approach is undertaken to investigate the micromechanics of wear of nuclear graphite IG-11. The effect of the height of surface asperities is introduced into the wear formula, and a correlation between dimensionless wear coefcient and surface asperity height is obtained. We learned that the likelihood of chip/particulate formation that is the major source of wear in metals is not dominant for porous, carbon-chain materials such as nuclear grade graphite. The constants in the proposed analytical wear model are obtained by tting with the computational results. Then, the model is used to predict experimental data. In this way, the accuracy of the proposed model is assessed. The computational results revealed that there is approximately 20% difference between the results of

References
ABAQUS Finite Element Analysis, SIMULIA, Dassault Systmes, http://www.simulia. com/support/documentation.html (visited May 2011July 2012). Archard, J.F., Hirst, W., 1956. The wear of materials under unlubricated conditions. Proc. R. Soc. A 236, 397410. Barlow, J., 1976. Optimal stress locations in nite element models. Int. J. Numer. Methods Eng. 10, 243251. Blau, P.J., 1992. ASM Handbook. Vol. 18: Friction, Lubrication, and Wear Technology. ASM International, Niagara Falls, NY, USA. Cogliati, J.J., Ougouag, A., 2008. Pebble bed reactor dust production model. In: Proceedings of the 4th International Topical Meeting on High Temperature Reactor Technology, Washington, DC, USA. Cogliati, J.J., Ougouag, A.M., 2010. Dust Production Model for HTR-10. Idaho National Laboratory, Report to Department of Energy. Cogliati, J.J., Ougouag, A.M., Ortensi, J., 2011. Survey of dust production in pebble bed reactor cores. Nucl. Eng. Des. 241 (6), 23642369. Czichos, H., 1978. Tribology: A System Approach to the Science and Technology of Friction, Lubrication and Wear. Elsevier Scientic Publishing Company, Bedford, NS, Canada. IAEA, 2003. Evaluation of High Temperature Gas Cooled Reactor Performance: Benchmark Analysis Related to Initial Testing of the HTTR and HTR-10, IAEATECDOC-1382. Irwin, G., 1957. Analysis of stresses and strains near the end of a crack traversing a plate. J. Appl. Mech. 24, 361364. Johnson, K.L., 1987. Contact Mechanics. Cambridge University Press, Fairford, GLO, UK. Johnson, G., 2012. Experimental study of graphitegraphite and graphitesteel wear in spherical contact in a pressurized inert atmosphere at elevated temperatures. University of Idaho, Idaho Falls, Idaho (Masters thesis). Kissane, M.P., 2009. A review of radionuclide in the primary system of a very-hightemperaure reactor. Nucl. Eng. Des. 239, 20763091.

508

M. Rostamian et al. / Nuclear Engineering and Design 263 (2013) 500508 U.S. Nuclear Regulatory Commission Regulation, 2008. Next Generation Nuclear Plant Phenomena Identication and Ranking Tables (PIRTs), NUREG/CR-6944 (13). Fission-Product Transport and Dose PIRTs. Xiaowei, L., Suyaun, Y., Zhen-sheng, Z., Shu-yan, H., 2005a. Estimation of graphite dust quantity and size distribution of graphite particle in HTR-10. Nucl. Power Eng. 26, 02580926. Xiaowei, L., Suyuan, Y., Xuanyu, S., Shuyan, H., 2005b. Temperature effect on IG-11 graphite wear performance. Nucl. Eng. Des. 235, 22612274. Yokoyama, T., Nakai, K., Futakawa, M., 2008. Compressive stressstrain characteristics of nuclear-grade graphite IG-11: effect of specimen size and strain rate. J. Jpn. Soc. Nucl. 7 (1), 6673.

Pook, L.P., 2007. Metal Fatigue, Solid Mechanics and its Applications, vol. 145. Springer, Guernsey, GY, UK. Rostamian, M., Arifeen, Sh., Potirniche, P.G., Tokuhiro, A., 2010. Initial analysis of pebble contact in pebble bed reactors. Trans. Am. Nucl. Soc. 103, 10281031. Rostamian, M., Arifeen, Sh., Potirniche, P.G., Tokuhiro, A., 2011. Initial prediction of dust production in pebble bed reactors. J. Mech. Sci. 2, 189195. Rostamian, M., Potirniche, P.G., Cogliati, J.J., Ougouag, A.M., Tokuhiro, A., 2012. Computational prediction of dust production in pebble bed reactors. Nucl. Eng. Des. 243, 3340.

You might also like