You are on page 1of 24

REMODELING OF FEMORAL STEM

BACKGROUND FEMUR The femur is the longest and strongest bone in the skeleton, is almost perfectly cylindrical in the greater part of its extent. In the erect posture it is not vertical, being separated above from its fellow by a considerable interval, which corresponds to the breadth of the pelvis, but inclining gradually downward and medialward, so as to approach its fellow toward its lower part, for the purpose of bringing the kneejoint near the line of gravity of the body. The degree of this inclination varies in different persons, and is greater in the female than in the male, on account of the greater breadth of the pelvis. Fractures A femoral fracture that involves the femoral head, femoral neck or the shaft of the femur immediately below the lesser trochanter may be classified as a hip fracture, especially when associated with osteoporosis. FEMORAL STEM The femoral stem component replaces a large portion of bone in the femur, and this is therefore the loadbearing part of the implant. To bear this load, it must have a Youngs Modulus comparable to that of cortical bone. If the implant is not as stiff as bone, then the remaining bone surrounding the implant will be put under increased stress. If it is stiffer than bone, then a phenomenon known as stress shielding will occur. DESIGN OF FEMORAL STEM

Design of the femoral stem is an important issue in the field of total hip arthroplasty, but design is just one component in the success or failure of the operation. Other components are surgical technique, cement technique or press-fit technique, bone quality, as well as patient related factors. The quality of design may not also be matched with quality of manufacturing and machining of the stem. The ultimate outcome of the arthroplasty obviously depends also on a matching acetabular component. Currently the femoral stem revision rate at 10-15 years is reported to be between 0% and 4.8% and does not correlate well with the radiographic stem loosening. Femoral stem design options are related to whether the stem is curved or straight, the presence or absence of collar support on the calcar, the stem cross section, the stem offset, the surface finish, as well as the value of stem modularity and some metallurgical issues.

Stem Offset?

The offset is the transverse distance between the centre of the head and the vertical line representing mid-stem or mid-femur (fig.9). Variability of offset helps to replicate the anatomy by insuring proper soft tissue tension (fig.10) which balances the hip bearings. Although a high offset stem relatively increases its bending moment (fig.11), various reports show that a high offset does not increase cement strain on medial cement mantle.

(fig.9) Stem offset is the distance between the head centre and vertical line representing the mid-stem.

(fig.10) The offset of the stem helps to replicate normal soft tissue tension.

(fig.11) High offset stem resulted in lateralising the femur by 4.69 mm, 8 years follow up x-ray.

A preoperative plan using a template (fig.12) of different offset stems help the surgeon to see which offset is likely to replicate the soft tissue anatomy.

(fig.12) X-ray template help in measuring the right offset for each patient.

Surface Finish?

How smooth should be the surface of the stem! Is a feature of great variation as it comes in five different ranges? Any surface will show peaks and valleys when examined by scanning electron microscopy (fig.13), the average between Peak and Valley is known as the Roughness Average (Ra); according to Ra the surface finish (fig.14) of femoral stems may be classified as:

1. 2. 3. 4. 5.

Highly polished Satin Matt Rough Textured

(fig.13) Metal surface, as seen by SEM, showing peaks and valleys.

(fig.14) Different stem finishing levels, polished (a), satin (b), Matt (c), rough (d), and textured (e) and (f).

A polished surface will show less fixation strength to cement, to the contrary of rougher surfaces which show greater fixation strength to cement.

Debonding is the loss of fixation between metal and cement. When debonding happens rough surfaces behave badly, as it will abrade the adjacent cement and will cause microfractures in cement mantle, ultimately leading to loss of fixation. This may lead also to the release into the effective joint space of abrasive wear debris from cement and metal, which when ground inside the bearings will act by 3rd body wear mechanism to release submicron poly wear particles initiating the process of osteolysis.

Surface Features?

These are any irregularities present on the stem surface apart from its finish discussed above, like flanges, serrations, centralizers, pre-coated beads and knobsetc..

The only surface features that may be beneficial are flanges and centralizers. Flanges are a part of the stem popularized in later Charnley design stems (fig.15) to help pressurize the cement as the proximal stem part is pushed into the femur.

(fig.15) The flanged design followed the roundback design in the Charnley stem series.

The stem centralizer (fig.16) is also beneficial as it prevents the stem from deviating in the canal, insuring even cement mantle and perhaps preventing an unwanted varus position of the stem. Non end-bearing centralizers may prevent cement fracture below the stem when subsidence occurs.

(fig.16) Stem centraliser insures a regular cement mantle and a centrally located stem.

Pre-coating with PMMA was a good idea assuming better cement bonding to the PMMA precoat as compared to metal. This did not seem to work, as there were reports by Mohler, 1995 of early femoral loosening in 2-10 years, other reported 15% stem failure rate over 6 years due to poor cement mantle and centralization.

Modularity?

Modularity helps intra-operative adjustment of components, most designs allow neck length (fig.17) and head size modularity, and a select few allow modularity in anteversion and CCD angle.

(fig.17) Modular neck length, the short, standard and long heads can vary the neck length, allowing adjustments during surgery.

The questions of increased wear and corrosion due to micromotion between the different pieces of the modular stem remain to be proven to assume a clinical disadvantage to these designs, however; the clinical problems of impingement / dislocation (e.g. by using a skirted extra-long head, or a very short head on a broad conical neck, fig.18) and of undue lengthening (fig.19) fall under the technical control of the surgeon, who must be aware of design and limitations of the stem he is implanting.

(fig.18) Using a short head, with a combination of conical neck and an antiluxation acetabular rim resulted in impingement during external rotation with slight abduction.

(fig.19) Leg lengthening due to the use of extra-long head

The modular stem costs more than the monoblock sibling, and adds to the logistics of the hospital creating more stock control overload on the administrator.

CHECK FROM DESIGN CRETERIA


Metallurgical Issues The current concept in hip arthroplasty prefers Cobalt-Chrome or Stainless steel for the cemented stems and Titanium for the cementless. Other ideas are also available; but the majority of surgeons world wide support this current concept. The Scope of Stem Design

This presentation stressed mainly on the standard cemented stem, but the scope of stem design is much larger, the cementelss stem may share many of the above points of discussion apart from those related to cement mantle and bonding. The surface of the cementless stem and its coating may warrant a separate article. The recent evolution of special stems used in femoral reconstruction and revisions is also not covered in this article, the author believe that these are better understood when discussed among topics related to complex femoral reconstruction and revision arthroplasty
PURPOSE OF FEMORAL STEM REMODELING A hip replacement with a femoral stem produces an effect on the bone called adaptive remodeling, attributable to mechanical and biological factors. All of the prostheses designs try to achieve an optimal load transfer in order to avoid stress-shielding, which produces an osteopenia. INTRODUCTION The implantation of a cemented or cementless femoral stem implies an important change in the physiological load distribution. The bone reacts to the new situation, in accordance with Wolff 's law, undergoing a process of adaptive remodelling [1], related to both mechanical and biological factors, being the most important the initial bone mass [2]. Achieving good primary xation is of crucial importance in cementless hip arthroplasty to ensure good short-term and long-term results. Lack of primary stability leads to thigh pain and eventual loosening of the prosthesis because of a continuous disruption of the bone formation process around the implant (Kim et al., 2003; Knight et al., 1998;Mont and Hungerford, 1997; Petersilge et al., 1997). The stability, or the lack of it, is commonly measured as the amount of relative motion at the interface between the bone and the stem under physiological load. Large interfacial relative movements reduce the chance of osseointegration, and cause the formation of a brous tissue layer at the boneimplant interface (Pilliar et al.,1986), which may eventually lead to loosening and failure of the arthroplasty. The threshold value of micromotion, above which a brous tissue layer forms, has been studied in both animals and humans. In a review of dental implants in animals, a threshold micromotion value between 50 and 150 mm was found (Szmukler-Moncler et al., 1998). A similar range of values was reported for orthopaedic implants in humans. In a retrieval study of cementless femoral components, Engh et al. (1992) found indications that micromotions less than 40 mm had resulted in osseointegration while micro-motions of 150 mm had caused the interposition of a brous tissue layer at the stembone interface. It can be concluded from these reports that the value of micromotion, above which osseointegration is disrupted, ranges from 50 to 150 mm, possibly skewed towards the lower end of this range. While many believe a sufficiently high interference t is essential to achieve good primary stability, it is also clear that introducing a interference fit has caused a clinically significant increase in intra-operative femoral canal fractures (Cameron, 2004; Meek et al., 2004), an effect which has also been demonstrated during in vitro testing (Jastiet al., 1993; Monti et al., 2001). The appropriate range of interference t that ensures primary stability without risking femoral fracture is not well understood.

There are in principle two parts to this study. In order to get a rough idea of the interference t introduced using current surgical practise, in the rst part of this study, nite element predictions were correlated with in vitro micromotion measurements. The aim of this was to enable back calculation of the real interference t introduced by the surgeon during the in vitro experiment. In the second part of the study, the effect of a range of interference ts on micromotion predictions was investigated using nite element models of a more physiologically realistic loading scenario than was possible during the rst part of the study. METHODOLOGY In the first part of the study, the finite element models were based on CT scans from the specific bones used in the experiment. In the second part of the study, the CT scans from the visible human dataset were used. Also in the first part of the study, the purpose was simply to compare finite element predictions and experiments and to simplify the experiments, a simple load configuration was chosen. In the second part of the study, physiological loads including muscle loads were used. In vitro experimental set-up The experiment was designed for direct comparison of micromotion values between experiment and FE analyses. Four cadaver femurs and Alloclassic (Zimmer GmbH, Winterthur, Switzerland) hip stems were used, and two points, one in the proximal part and another in the distal part of the stem (Fig. 1), were chosen for micromotion measurement. In order to avoid damaging the stembone interface during drilling action, the two points on the implant were drilled before implantation. A guide jig ensured that the bone, subsequent to stem insertion, was drilled in the position matching these same two points on the stem. Finally, steel pegs were glued into the holes in the stem and protruding through the bone (Fig. 1, right). A linear variable differential transducer (LVDT Model DFg5, DC Miniature series, Solartron Metrology, UK), was rigidly xed to the outside of the femur (Fig. 1, right). The connecting rod of the LVDT core rested on the free-end of the steel peg. When the implant was loaded, the implant and hence the peg moved relative to the bone and the LVDT measured the axial movement of the peg relative to the transducer, thus providing an estimate of the relative axial movement between bone and stem. Implantation was carried out by an experienced orthopaedic surgeon (D.L.). The neck of the femur was rst resected, and the femur was then reamed with rm impaction using a series of reamers to open the canal. A femoral stem was then implanted in the femur.

Fig. 1. The jig used to position the holes in the bone and the pegs in the implant, respectively (left). The implantbone specimen with LVDT attached to the femur loaded in compression in the mechanical testing machine (right). The femur was sectioned 250mm distal to the lesser trochanter and its distal end xed inside a cylindrical metal container using polymethyl- methacrylate (PMMA). These were then placed onto the table bed of a universal materials testing machine (Instron 5565, Instron Corp., Canton, MA). The specimen was adjusted so that the long axis of the stem was coaxial to the direction of loading. A cyclical axial compression load of 02 kN and triangular waveform was applied to the shoulder of the stem for 50 cycles at a rate of 1 kN/min using a 5 kN load cell. Micromotion readings via the LVDT were taken manually at maximum load of 2 kN and when fully unloaded at each cycle. Finite element methodology for correlation study A 3D model of a hip stem (Alloclassic, Zimmer GmbH) was constructed from CAD les received from the manufacturer (Fig. 2).

Fig. 2. The hip stem used in the study indicating the FE mesh used (left) and the implant inserted in the femur (right). In the correlation part of the study, the nite element model needs to be as accurate a representation of the experimental set-up as possible. Hence, the FE simulations of this part of the study were based on CT scans of the specic bones used in the experiments. There were two sets of scans: one scan prior to inserting the implant in the femur and a subsequent scan after implantation. The rst set of scans was used to derive bone geometry and material properties from the Houndseld units of the scan, while the second set of scans was used to ensure that the implant position and orientation in the FE model precisely matched the implant position within the femur in the experiment. The reason for this two-step procedure is that it would be inappropriate to use the CT datasets from the implanted femur for bone property assignment due to artefacts in these datasets caused by the metal stem. The construction of 3D models of the hip was done using AMIRA software (Mercury Computer Systems, Inc., San Diego, CA). Segmentation was compiled automatically using the softwares marching cubes algorithm which generates a 3D triangular surface mesh. The completed model was then converted to solid linear tetrahedral elements using Marc.Mentat (MSC.Software, Santa Ana, CA) software. The mesh was inspected to ensure it was reasonably shaped throughout. The Marc nite element software package was used in this study. Material properties for the bone were assigned based on the grey-scale value of the CT images on an element-by-element basis. The grey-level of the CT images was related to the apparent density using a

linear correlation (Cann and Genant, 1980; McBroom et al., 1985). This allowed for the transformation of the spatial radiological description into the description of bone density. The modulus of elasticity of individual elements was then calculated from the assigned apparent densities using the cubic relationship proposed by Carter and Hayes (1977). The material properties were assumed to be linear elastic and isotropic with Poissons ratio set to 0.35. The FE model was loaded at the centre of the shoulder of the stem with 2 kN, the stem being coaxial to the direction of loading, hence, matching the loading conguration in the experiment. Mesh convergence is a standard issue in any nite element analysis and in a contact analysis, there are many other numerical parameters that affect the predicted micromotions. The default contact strategy in Marc is a direct constraint algorithm (MSC.Marc-Manual, 2004) which most importantly requires the input of a contact zone size (CZS). Furthermore, Bernakiewicz and Viceconti (2002) described the importance of the convergence tolerance (CTol) in non-linear analyses. They also suggested that the appropriate parameter settings should be such that the resultant change in predicted micromotion between models with different parameter settings should be small relative to 150 mm. A sequential sensitivity analysis involving mesh density, CZS and CTol was carried out and a model with 12,078 nodes, CZS 0.025mm and CTol 1% was found to be sufcient for an accurate solution. We also chose a Coulomb friction model which in Marc requires theinput of the friction coefcient (m) as well as a parameter (SL). The Marc software has introduced the parameter SL, which describes a smoothing of the step-function of the Coulomb model, only in order to deal with an otherwise numerically difcult to handle discontinuity. However, not only does this parameter dramatically affect the predicted micromotion (Fig. 3) it also has an important physical interpretation. Shirazi-Adl et al. (1993) showed that the boneimplant interface friction curve is highly non-linear, exhibiting micromotion on the order of 150 mm (that is in the order of the critical level for osseointegration) before the slip load predicted by the Coulomb model is reached. The implication of Shirazi-Adl et al.s work is that adopting the ideal Coulomb model is inadequate. However, the SL parameter can be interpreted and used to represent this non-linear behaviour.

Fig. 3. Contour plots of micromotion over the surface of the Alloclassic stem under stairclimbing loads and for different values of the SL parameter (SL describes the non-linear friction characteristics of the interface). To establish the appropriate setting of the SL parameter, we simulated Shirazi-Adl et al.s relatively simple experiment consisting of a bone cube exposed to normal and tangential loads moving on a metal plate. In Fig. 4 is shown Shirazi-Adls experimental curve of tangential load versus tangential displacement. The tangential load that would initiate slip according to the Coulomb model is 30.6. The nite element predicted curves for various settings of SL is also shown and a setting of SL 0.1 predicts the experimental curve well. Hence, in the rest of this study, this setting was used.

Fig. 4. Tangential load versus tangential displacement of bone cube sliding on metal plate. The nite element predicted non-linear friction behaviour for different levels of the parameter SL is shown as well as the experimental curve reported by Shirazi-Adl et al. (1993). The critical value at which sliding would initiate according to an ideal Coulomb friction model is also indicated. The effect of friction coefcient on micromotion is relatively minor for friction coefcients higher than 0.15 (Kuiper and Huiskes, 1996). Viceconti et al. (2000) found that a friction coefcient between 0.2 and 0.5 led to the best correlation with experiments. Rancourt et al. (1990) measured friction coefcients experimentally and found a coefcient of 0.4. Based on these previous studies, a friction coefcient of 0.4 was used in this study. The objective of this study was to estimate the effective interference t. Hence, we varied the interference t in the nite element models. The predictions were then compared to the experimentally measured values to estimate which level of interference best matched the experiment.

Finite element methodology for parametric study on the effect of interference t In this part of the study, geometry and material characterisation was based on the CT scans available from the visible human project (VHP) dataset. The hip stem model was aligned inside the femur according to the recommendations of the manufacturer (Alloclassic surgical technique, Zimmer Ltd., Warsaw, IN) and an experienced surgeon (D.L.) inspected the resulting congurations and considered them appropriate. The models were restrained distally and loaded with physiological stair-climbing loads including all relevant muscle forces. Stair-climbing loads were applied as this loading scenario has been shown to be more critical than other activities (Kassi et al., 2005). Similarly, Kassi et al. (2005) showed that it is essential to include muscle loads although this issue has been debated (Cristofolini and Viceconti, 2006). This load conguration was based on the extensive work by Bergmann (2001) and Heller et al. (2005) in which load directions and muscle attachments are described. The magnitudes of the loads in percentage body weight are shown in Table 1 and a body weight of 82 kg was used. All other aspects of the model were as described in Section 2.2. RESULTS In vitro micromotion measurements and correlation with model predictions During the experiments, the stem initially subsided into the bone but after a sufcient number of cycles the stem settled. Even in this relatively stable state, there continued to be low levels of reversible motion at the stembone interface as a result of the continued loadunload cycle. It is high levels of this continued disruption of the interface that is thought to prevent osteogenic cells from bonding to the surface of the stem (Pilliar et al., 1986). Hence, in terms of evaluating the ingrowth potential of an arthroplasty, it is the reversible micromotion rather than subsidence which is the relevant constituent of the overall relative motion between bone and implant. The reversible micromotion during a load cycle was estimated as the difference between the total micromotion measured at maximum load and the total micromotion when the specimen was unloaded. For the remainder of this study, we will refer to this quantity as reversible micromotion or just micromotion.

In Fig. 5 is shown the experimentally determined reversible micromotion for each of the four specimens plotted against load cycles for the distal and proximal parts, top and bottom, respectively. During the initial cycles, this micromotion was high but then stabilised at a lower value. Relatively high levels of micromotion during the rst few times a patient exposes a joint to loading, are probably not critical. It is the long term or stabilised value of reversible micromotion that will continue to disrupt the implantbone interface and prevent osseointegration. Hence, it is the converged values of Fig. 5 which are relevant. Based on the data of Fig. 5, the converged average value in the distal and proximal regions were 1872 and 1975 mm, respectively.

Fig. 5. Distal micromotion (top) and proximal micromotion (bottom) results from the experiment.

The results of the FE analyses using different levels of interference t and simulating the experiment are shown in Fig. 6. The gure shows that with just 1 mm of interference, the level of micromotion is predicted to be in the range of 2030 mm. With 2 mm of interference, this drops to 1020 mm. Comparing this to the experimental values of 18 and 19 mm also shown in the gure, this implies that the interference t introduced by the surgeon is only 1 or 2 mm.

Fig. 6. Contour plots of micromotion over the surface of the stem under an axial load of 2 kN, using interference ts of (from left to right) 0, 1, 2 and 5 mm, respectively. The experimentally determined proximal and distal micromotion is also indicated. This seems perhaps unrealistically low. Shultz et al. (2006) considered an interference t of 100 mm to cause bone interface damage and reported this level of interference as a threshold value. Therefore, we included an interference t of 100 mm in one of the nite element models and inspected the resulting tensile hoop stresses (Fig. 7). This model was not exposed to any other loads. As can be seen from the gure, interference induced hoop stresses are on the order of 50MPa on the surface of the bone (internally the stresses are somewhat higher).Comparing this stress level with the transverse tensile strength of cortical bone of approximately 50MPa (Reilly and Burstein, 1975), it would seem that 100 mm represents the critical level of effective interference t above which the femoral canal will fracture. The location of high hoop stresses towards the distal end of the implant seen in Fig. 7 also matches the location of 77% of intra-operative fractures (Meek et al., 2004).

Fig. 7. Hoop stresses in femoral bone caused by an interference t of 100 mm. No external loads are applied in this model. Considering that femoral canal fractures are not infrequently occurring intra-operatively (Cameron, 2004; Meek et al., 2004), it would seem that surgeons are introducing close to the critical level of interference t of 100 mm. Assuming that surgeons are able to control the insertion process within a factor of 2, perhaps a realistic range of interference t can be argued to be in the range of 50100 mm. In summary, this rst part of the study indicates that the range of realistic interference ts may be within a range of very low levels (just a few microns) and up to 100 mm. The effect of interference t on micromotion Fig. 8 shows contour plots of predicted micromotion over the stem surface under stairclimbing loads and for four different levels of interference t. Fig. 9 shows the change in micromotion with levels of interference t for the two points labelled P (proximal) and D (distal) shown on the left model of Fig. 8. Also in Fig. 9 is indicated, by the grey-coloured region, the threshold range of micromotion above which soft tissue formation will be predicted and below which osseointegration would be expected. From these two gures, it is clear that the interference t had a very large effect on micromotion predictions. In the case of no interference t, the entire surface area of the implant was in or above the grey area indicating that the primary stability of the implant is at risk. In contrast, with 50 mm of interference, all but the most proximal part of the implant was predicted to osseointegrate. Interestingly, increasing the level of interference beyond 50 mm had negligible effect. Also, it is clear that the effect of the interference t was most dramatic at low levels of interference. Including just 5 mm of interference causes almost a 50% reduction in micromotion and including more interference only has a relatively small effect.

Fig. 8. Contour plots of micromotion over the surface of the stem under stairclimbing loads and with interference ts of (from left to right) 0, 5, 25 and 50 mm, respectively.

Fig. 9. Micromotion at points P (proximal) and D (distal) as a function of the level of interference t. Locations of point P and D are shown in Fig. 8 (left). The grey area indicates the range of the critical micromotion threshold. Above this level, brous tissue formation would be expected; below, osseointegration is anticipated.

CONCLUSION
This study has shown that modeling the interference fit characteristic of hip stems is crucial for quantitative predictions of micro-motion. Ignoring the interference fit will probably lead to an under estimation of the stability of the stem. In contrast, ignoring the non-linear friction behavior reported by Shirazi-Adl et al. (1993) and reproduced in Fig. 4, will probably to lead to too optimistic predictions of stem stability. The magnitude of interference fit is fundamentally unknown and may be the reason most previous works have omitted this parameter from their finite element analyses. Indeed, during this study it became clear just how difficult it is to estimate this parameter. Nevertheless, this study demonstrates the importance of the interference fit as including only a small level of interference changed the evaluation of the investigated stem from that of an unstable stem to that of a stable stem. Our predictions showed high levels of micro-motion distally and proximally while micro-motion at the stem midsection was lower (Fig. 8, left). This is qualitatively consistent with the finite element predictions by Keaveny and Bartel (1993). Keaveny and Bartel did not include an interference fit and predicted very high absolute values of micromotion (0550 mm). Keaveny and Bartel simulated a cylindrical stem which is likely to be less resistant to torsional loads and that may explain the higher levels of micromotion as compared to our results. Viceconti et al. (2000) did simulate a press-fit although it is not possible to quantify this press-fit in a manner that allows a direct correlation with our results. Viceconti et al. predicted micromotions ranging from 17 to 49 mm across the surface of the implant which is reasonably consistent with our results simulating an interference fit of 25 mm (Fig. 8). The results of Fig. 6 indicate that surgeons introduce very low interference fits, on the order of 1 2 mm. Apart from any aspects of the model that may cause inaccurate predictions, it is of course also possible that the experimental results are inaccurate. Notably, our experiment, like the vast majority of other experimental micromotion studies, does not measure the actual interface micromotion but instead measures the motion between the LVDT fixation point on the bone and the point of the peg insertion on the implant. The motion measured, therefore, includes other flexibilities such as bone deformation and will tend to overestimate micromotion (Bu hler et al., 1997). If these flexibilities are substantial compared to the true interface micromotion, it would cause our methodology to predict very small levels of interference which is of course what seems to be the case. In connection with Fig. 7, we proposed that surgeons are in fact more likely to introduce interference fits of 50100 mm. Shultz et al. (2006) predicted that with an interference fit of 100 mm, the hoop stresses in the bone would visco-elastically relax by approximately 50%. In other words, if a surgeon introduces an interference fit of 100 mm, this would relax and represent an effective interference of 50 mm. Shultz reported that interference fits lower than 100 mm would relax less than 50%. Therefore, even if a surgeon only achieves the lower range of the50100 mm interference, we have estimated, there should be at least 25 mm of effective interference left after relaxation, well above the 1 2 mm estimated from the experiment. We have no evidence to explain the small levels of interference fit predicted from the experiments but we are inclined to believe that the experiment overestimated the micromotion, for the reasons noted above. We have assumed a uniform interference fit over the entire surface of the implant. Accordingly, the pressfit (pressure) varied considerably from the proximal cancellous femur to the cortical distal femur as modelled through the variation in the local Youngs modulus of the bone adjacent to the implant. This variation in press-fit between the proximal and distal region is undoubtedly qualitatively correct. However, our study was not set up to investigate variation in interference fit. This was not included due to the practical difficulty in quantifying the variation and generalising such variation that is likely to vary between implants. It is also probable, given the very small interferences calculated, that surgeons cannot create implant cavities with uniform interference across the interface area, so that clinical cases would include variations from the micromotions predicted. The effects of a more realistic scenario are not yet known.

The results of this study support the suggestion made earlier (Shirazi-Adl et al., 1994) that the cavity that is created in the femur is larger than is indicated by the nominal interference of 0.3 0.5mm (Otani et al., 1995;Ramamurti et al., 1997); such a large interference would cause the femur to fracture, according to our results. Perhaps the most important result of the study and the result with direct clinical relevance relates to Figs. 7 and 9. Fig. 7 predicts that surgery is safe against femoral canal fracture at interference fits lower than 100 mm. Fig. 9 predicts that the stem would osseointegrate at interference levels of 50 mm. Therefore, the recommendation is for the surgeon to err on the side of a low interference fit during surgery as only 50 mm is enough to achieve stability and provides a safety factor of 2 against femoral canal fracture. If considering a stem likely to be successful as long as just the distal part of the stem (embedded in the strong cortical bone) osseointegrates, Fig. 9 indicates that just 10 mm of interference fit is necessary for stability and provides a safety factor of 10 against femoral canal fracture. Of course, our computational predictions should be further investigated before being applied in clinical practise. It is likely, that stems with different geometry or material will behave differently. The Alloclassic stem in this study, for example, has a rectangular cross-section, which might be advantageous in resisting torsional loading during the stairclimbing simulated. Nevertheless, the predictions clearly indicate a recommendation to modify surgical practise thereby reducing or even eliminating the 7% intra-operative femoral canal fractures during primary hip surgery reported by Cameron, (2004) and the 650% fracture rates reported by Meek et al. (2004) in connection with revision hip surgery.

You might also like