The Orifice Pins
Geometrical Model
Although four versions of the orifice pin were created by Gimeno et al. (6), we consider here only the 2-orifice and 8-orifice pins, representing the extreme cases. The orifice pins (Fig. 2) comprise a central drug-filled hollow core (\( {\Omega}_3 \)) connected to orifices (\( {\Omega}_2 \)) drilled through the cylindrical wall of the pins. Following correspondence with the authors, it was revealed that the release medium and orifice pins were contained within 250 mL conical flasks, whose dimensions are available from the manufacturer (16). Although the release medium vessel is a conical flask, only the volume occupied by the release medium is required for the computational modelling and so the neck of the conical flask is not included in the computational geometry. Moreover, the authors confirmed that the pins were suspended in the release media. The geometry of each pin and release container was constructed in COMSOL Multiphysics®, version 5.3a, where the default Cartesian (x, y, z) co-ordinate system was employed. In Fig. 2 we display the geometries associated with the 8-orifice pin, highlighting the notation used to define each domain and boundary.
Fluid dynamics
While the release medium used in (6) is known to be Simulated Body Fluid (SBF), we have been unable to identify the fluid properties (e.g. density and kinematic viscosity) of SBF in the literature. For the purposes of this study, we assume that SBF has similar properties to those of water and, therefore, describe the fluid dynamics in the container and within the pin using the incompressible Navier-Stokes equations (1) supplemented with appropriate boundary and initial conditions (2-4):
$$ {\displaystyle \begin{array}{lcrr}\frac{\partial \boldsymbol{u}}{\partial t}& =& \nu {\nabla}^2\boldsymbol{u}-\left(\boldsymbol{u}\cdot \nabla \right)\boldsymbol{u}-\frac{1}{\rho}\nabla p,\kern1em \nabla \cdot \boldsymbol{u}=0,& \\ {}& & \mathrm{in}\kern2.77695pt {\Omega}_1,{\Omega}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t>0,\end{array}} $$
(1)
$$ \boldsymbol{u}=\left(-r\sin \left(\theta \right),r\cos \left(\theta \right),0\right)\omega, \kern1em \mathrm{on}\kern2.77695pt {\Gamma}_1,\kern1em t>0, $$
(2)
$$ \boldsymbol{u}=\boldsymbol{0},\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_2,\kern1em t>0, $$
(3)
$$ \boldsymbol{u}=\boldsymbol{0},\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,{\Omega}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t=0, $$
(4)
where u denotes the fluid velocity field, assumed to be zero initially (4), \( \rho \) is the fluid density, \( p \) is the pressure and \( \nu \) is the kinematic viscosity of the fluid. On the boundary Γ1, a moving wall boundary condition (2) is imposed. For ease of implementation, a cylindrical co-ordinate system was employed for this boundary condition, in which \( \omega \) is the magnitude of the angular velocity of the rotating wall in rad/s. This condition ensures that the fluid at this boundary takes the wall’s angular velocity: this is the driving force behind the fluid flow. A no-slip/no-penetration boundary condition (3) is applied to the collection of boundaries that comprise the body of the pin, denoted by Γ2. We note that this model inherently assumes that the submerged pin is filled with fluid at t = 0. In reality, the SBF will take some finite time to infiltrate the pin and will depend on many factors including the diameter of the orifices and the structure of the (porous) dry drug core. In cases where the timescale associated with drug dissolution is significant, this model may represent an under-estimation of the drug release time.
Drug transport
Drug transport within the release medium may in principle be governed by advection as a result of fluid flow as well as diffusion as a result of random molecular motion. Additionally, when the drug is loaded in a dry solid form, as is the case in Gimeno et al. (6), the rate of dissolution of drug may be a key driver of drug release. Indeed, (6) considered in their study two commercially available drug products with differing solubilities (Cefazolin Sodium and Linezolid, with solubilities in water of 50 mg/mL and 3 mg/mL, respectively), noting significant differences between the drug release profiles. Thus, to enable the possibility of each of these mechanisms being equally important, we propose a dissolution-diffusion-advection model to describe drug transport within the hollow interior of the pin:
$$ \frac{\partial b}{\partial t}=-\beta {b}^{2/3}\left(S-c\right),\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t>0, $$
(5)
$$ \frac{\partial c}{\partial t}={D}_f{\nabla}^2c-\boldsymbol{u}\cdot \nabla c+\beta {b}^{2/3}\left(S-c\right),\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t>0, $$
(6)
where \( b\left(x,y,z,t\right) \) and \( c\left(x,y,z,t\right) \) are the concentrations of undissolved and dissolved drug, respectively, \( S \) is the solubility of the drug in the release medium, \( \beta \) is the drug dissolution rate, and \( {D}_f \) is the isotropic free-diffusion coefficient of the drug in the release medium. This nonlinear dissolution model was originally proposed by (17) and represents a modification of the classical Noyes-Whitney model (18) allowing for the possibility of the dissolution rate being a function of the surface area of spherical dissolving drug particles. The model (5-6) assumes that the orifice pins are rapidly infiltrated by the release medium fluid and that the drug is fully wetted. This enables the dissolution process to begin throughout the drug core immediately. In reality, dissolution is a complex process consisting of several steps. There are more complex models in the literature (19) which account for each individual step, highlighting the importance of the initial wetting stage. The equation proposed to model mass transport in the bulk fluid domain and in the orifices is:
$$ \frac{\partial c}{\partial t}={D}_f{\nabla}^2c-\boldsymbol{u}\cdot \nabla c,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1\mathrm{and}{\Omega}_2,\kern1em t>0. $$
(7)
We impose the following boundary and initial conditions
$$ \hat{\boldsymbol{n}}\cdot \left(-{D}_f\nabla c+c\boldsymbol{u}\right)=0,\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_1,\kern1em t>0, $$
(8)
$$ -\hat{\boldsymbol{n}}\cdot {D}_f\nabla c=0,\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_2,\kern1em t>0, $$
$$ b={b}_0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t=0, $$
(9)
$$ c=0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,{\Omega}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t=0, $$
(10)
where \( \hat{\boldsymbol{n}} \) is the outward facing unit normal to applicable boundary surfaces and \( {b}_0 \) is the initial drug loading concentration.
The Porous Pin
Geometrical Model
The porous pin is comprised of two distinct regions: the inner drug core (Ω3) and the porous wall (Ω2) as shown in Fig. 3a. In the absence of any information to suggest otherwise, the container for the release medium is assumed to be a beaker of radius 30 mm (20). Since the volume of release medium is known to be 100 ml (5), the depth of the release medium may be easily calculated to provide as accurate a representation as possible for the model. The porous pin geometry (Fig. 3a) is then suspended in the centre of the beaker geometry as shown in Fig. 3b. This configuration was motivated through correspondence with the authors of the original work detailing the prototype pins (5). The geometry of the porous pin and release container were constructed in COMSOL Multiphysics®, version 5.3a, where the default Cartesian (x, y, z) co-ordinate system was employed. The notation used to define each domain and boundary is show in Fig. 3.
Fluid Dynamics
As with the orifice pins experiments, the release medium is SBF (5) with fluid dynamic properties assumed similar to water. The time-dependent incompressible Navier-Stokes equations (11,13) are used to model fluid flow in the bulk release medium and the inner drug core of the pin. However, these equations are not appropriate for fluid modelling in the porous wall of the pin (\( {\Omega}_2 \)), where we instead impose the time-dependent incompressible Brinkman equations (12,13). The Brinkman equations, preferred to Darcy’s law because of their more accurate approximation of bulk fluid-porous domain transitions (21,p. 16), introduce two additional parameters: \( \kappa \), the permeability of the porous wall of porosity \( \phi \). respectively. Furthermore, ν is the kinematic viscosity of the fluid (the ratio of dynamic viscosity to density). The fluid dynamics model is then given by:
$$ \frac{\partial \boldsymbol{u}}{\partial t}=\nu {\nabla}^2\boldsymbol{u}-\left(\boldsymbol{u}\cdot \nabla \right)\boldsymbol{u}-\frac{1}{\rho}\nabla p,\kern5pt \mathrm{in}\kern2.77695pt {\Omega}_1\mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t>0, $$
(11)
$$ \frac{1}{\phi}\frac{\partial \boldsymbol{u}}{\partial t}=\frac{\nu }{\phi }{\nabla}^2\boldsymbol{u}-\frac{1}{\phi^2}\left(\boldsymbol{u}\cdot \nabla \right)\boldsymbol{u}\kern5pt -\frac{1}{\rho}\nabla p-\frac{\nu }{\kappa}\boldsymbol{u},\mathrm{in}\kern2pt {\Omega}_2,\kern6pt t>0, $$
(12)
$$ \nabla \cdot \boldsymbol{u}=0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,{\Omega}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t>0. $$
(13)
These equations are supplemented with appropriate boundary and initial conditions (14-17):
$$ \boldsymbol{u}=\left(-r\sin \left(\theta \right),r\cos \left(\theta \right),0\right)\omega, \kern1em \mathrm{on}\kern2.77695pt {\Gamma}_1,\kern1em t>0, $$
(14)
$$ {\boldsymbol{u}}_{ns}={\boldsymbol{u}}_{br},\kern1em {p}_{ns}={p}_{br},\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Gamma}_3,\kern1em t>0, $$
(15)
$$ \boldsymbol{u}=\boldsymbol{0},\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_4,\kern1em t>0, $$
(16)
$$ \boldsymbol{u}=\boldsymbol{0},\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,{\Omega}_2\kern2.77695pt \mathrm{and}\kern2.77695pt {\Omega}_3,\kern1em t=0. $$
(17)
The definitions of \( \boldsymbol{u} \), \( \rho \), \( p \), ν and \( \omega \) are the same as with the orifice pins case. As before, a moving wall boundary condition (14) is used. Also, a no-slip/no-penetration boundary condition (16) is applied along with the assumption that the release medium is at rest initially (17). At the boundaries \( \boldsymbol{u} \), \( \rho \), \( p \), ν and \( \omega \)we impose continuity of velocity and pressure (15), where the subscripts \( ns \) and \( br \) indicate the velocity field and pressure associated with Navier-Stokes (\( ns \)) and Brinkman (\( br \)), respectively.
Drug Transport
The fluid dynamic equations (11-17) are coupled with a set of reaction-diffusion-advection equations (18-21). The drug concentration in the inner drug core (\( {\Omega}_3 \)), the porous wall of the pin (\( {\Omega}_2 \)) and the bulk release medium (\( {\Omega}_1 \)) are denoted \( {c}_d\left(x,y,z,t\right) \), \( {c}_p\left(x,y,z,t\right) \) and \( {c}_m\left(x,y,z,t\right) \), respectively. The proposed model is:
$$ \frac{\partial b}{\partial t}=-\beta {b}^{2/3}\left(S-c\right),\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t>0, $$
(18)
$$ {\displaystyle \begin{array}{lcrr}\frac{\partial {c}_d}{\partial t}& =& {D}_f{\nabla}^2{c}_d-\boldsymbol{u}\cdot \nabla {c}_d+\beta {b}^{2/3}\left(S-c\right),& \\ {}& & \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t>0,\end{array}} $$
(19)
$$ \phi \frac{\partial {c}_p}{\partial t}={D}_p{\nabla}^2{c}_p-\boldsymbol{u}\cdot \nabla {c}_p,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_2,\kern1em t>0, $$
(20)
$$ \frac{\partial {c}_m}{\partial t}={D}_f{\nabla}^2{c}_m-\boldsymbol{u}\cdot \nabla {c}_m,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,\kern1em t>0, $$
(21)
where \( {D}_p \) is the effective diffusion coefficient within the porous wall, calculated via \( {D}_p={\phi}_e{D}_f/\tau \)(22). The parameters ϕe and τ are the effective porosity of the porous wall and the tortuosity, respectively. These governing equations are supplemented with the boundary and initial conditions:
$$ {c}_d={c}_p,\kern1em \hat{\boldsymbol{n}}\cdot \left(-{D}_f\nabla {c}_d+{c}_d\boldsymbol{u}\right)=\hat{\boldsymbol{n}}\cdot \left(-{D}_p\nabla {c}_p+{c}_p\boldsymbol{u}\right),\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_3,\kern1em t>0, $$
(22)
$$ {c}_p={c}_m,\kern1em \hat{\boldsymbol{n}}\cdot \left(-{D}_p\nabla {c}_p+{c}_p\boldsymbol{u}\right)=\hat{\boldsymbol{n}}\cdot \left(-{D}_f\nabla {c}_m+{c}_m\boldsymbol{u}\right),\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_2,\kern1em t>0, $$
(23)
$$ \hat{\boldsymbol{n}}\cdot \left(-{D}_f\nabla {c}_m+{c}_m\boldsymbol{u}\right)=0,\kern1em \mathrm{on}\kern2.77695pt {\Gamma}_1,\kern1em t>0, $$
(24)
$$ {\displaystyle \begin{array}{lcrr}\hat{\boldsymbol{n}}\cdot {D}_f\nabla {c}_d& =& \hat{\boldsymbol{n}}\cdot {D}_p\nabla {c}_p=\hat{\boldsymbol{n}}\cdot {D}_f\nabla {c}_m=0,\kern1em & \\ {}& & \mathrm{on}\kern2.77695pt {\Gamma}_4,\kern1em t>0,\end{array}} $$
(25)
$$ b={b}_0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em t=0, $$
(26)
$$ {\displaystyle \begin{array}{lcrr}{c}_d& =& 0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_3,\kern1em {c}_p=0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_2,\kern1em & \\ {}{c}_m& =& 0,\kern1em \mathrm{in}\kern2.77695pt {\Omega}_1,\kern1em t=0.\end{array}} $$
(27)
Continuity of concentration and flux conditions (22–23) are applied to boundaries Γ2 and Γ3. On Γ1 and Γ4, zero-flux boundary condition (24–25) are applied to prevent drug from leaving the system, with the latter neglecting the advective component due to the zero-flux/zero-penetration conditions imposed on the fluid (16). Initially, there exists only undissolved drug (26) and no dissolved drug (27).
Summary of Investigated Scenarios
Since the focus is on establishing the influence of flow on drug release, we initially neglect dissolution and solve the corresponding 3D advection-diffusion models with all of the drug assumed to be in the dissolved phase initially at concentration \( {b}_0 \). This approximation is valid when the drug is either initially in a dissolved form, or is rapidly dissolved (e.g. high dissolution rate and/or solubility). We also consider the case of steady flow versus time-dependent flow. Once the effect of flow is established, we then reintroduce dissolution to the models. Given the geometry of the porous pin, we proceed to explore whether or not it is possible to exploit symmetry to simplify the model. Finally, when we have established the feasible simplifications to the model for each pin, we conduct a sensitivity analysis to explore the effect on drug release of varying the key model parameters. The key measure we use to compare the results is the drug release profile, defined as the mass of drug that has accumulated in the release medium (or released from the pins) at any time, t, normalised with respect to the initial drug loading.
Numerical Solution
The orifice and porous pin model equations (1-27) were nondimensionalised prior to solving numerically and applied to the geometries shown in Figs. 2 and 3, respectively. Additionally, a 2-orifice pin geometry was also considered, with the orifices located as shown in Fig. 1a. All spatial variables were scaled with the radius of the drug core, \( {L}_d \). The remaining scalings employed were:
$$ {c}^{\prime }=\frac{c}{S},\kern1em {b}^{\prime }=\frac{b}{b_0},\kern1em {t}^{\prime }=\frac{D_f}{L_d^2}t,\kern1em {\mathbf{u}}^{\prime }=\frac{L_d}{D_f}\mathbf{u},\kern1em {p}^{\prime }=\frac{L_d^2}{\rho {D}_f^2}p. $$
The models were solved using the commercial finite element method (FEM) software, COMSOL Multiphysics®, version 5.3a. There exist several numerical methods within the software, with the default methods for a given combination of physics selected by the software automatically. The particular methods considered for each study are described in the Results and Discussion Section. However, in all numerical studies, time-advancement was handled by the backward differentiation formula (BDF), with free time stepping. To aid in conservation of mass when the drug-transport equations had an advective component, the equations of each study were solved in conservative form. Preliminary modelling suggested that mass conservation could be influenced by the specific values of the parameters. Therefore, suitably dense meshes were constructed for use across all parameters considered. The meshes used in all studies were considered suitable if the error in mass conservation was less than 1%. For the models involving the Navier-Stokes equations, since the pressure is not defined anywhere in the system, a pressure point constraint was imposed by setting the pressure to zero at an arbitrary point in the pin domains to allow the numerical scheme to find a solution.
Parameter Values
A summary of the baseline parameter values used in this study is provided in Table 2. Several of the parameters have been taken or inferred from (5) and (6). The initial drug concentration and solubility values are based upon the antibiotics used in these studies: Cefazolin for the Orifice pins and Linezolid for the Porous Pin. The drug loading in the case of the orifice pins is explicitly stated (100 mg (6)), therefore, the initial drug concentration can be calculated. However, in the case of the porous pin, the authors state that the porous pins were loaded with 95 − 120 mg of the drug for the release experiments. For simplicity, we assume the drug loading is the same as in the case of the orifice pins (100 mg). We now discuss the remaining parameters.
Table 2 Baseline parameter values The free diffusion coefficient of drug in the release medium is assumed to be of O(10-9) m2/s with the upper limit value of 1×10-9 m2/s selected for the baseline case. The effective porosity of the porous wall, \( {\phi}_e \) is taken to be 90 % of the overall porosity, while the tortuosity, \( \tau \), is assumed to take the value 3, which is considered an average value of the typical range of tortuosities (22). We are unaware of literature estimates of the dissolution rate, \( \beta \), for these drugs. Therefore, we selected the baseline value such that the associated second Damköhler is of O(1) (23). The rate of stirring in the orifice pin experiments was quoted as 30 RPM, however, the rate of stirring in the porous pin experiments was not reported, therefore, we assume it to also be 30 RPM. The kinematic viscosity of the release medium (ν) was inferred from (24) while the permeability of the porous wall \( \kappa \) was derived from (25).
In order to assess the effect of varying the model parameters on the resulting drug release profile, a sensitivity analysis was conducted whereby several of the model parameters were varied from the baseline values.