Our product, BOROHMA, is a mosquito-repellent fragrance that prioritizes comfort without compromising effectiveness. Apart from experimental data, we also wanted to use drylab efforts to further validate the functionalities and market viability of our BOROHMA. To perform these validations, we developed three models:

  1. Production-Release Model

    To find the concentration of borneol released using enzyme kinetics and minicell diffusion to ensure an adequate concentration to repel mosquitoes.

  2. Stickiness Metric Framework

    To quantify the comfort of our repellent by modeling the viscosity over time of our base minicell solution and relating that to the stickiness coefficient.

  3. Market Adoption Model

    To predict how BOROHMA would perform on the market, being a unique, cross-market product that combines features of both repellents and fragrances.


Click the tabs to view each of our models!


Production-Release Model

Overview

To demonstrate the effectiveness of BOROHMA, we combined two models to simulate the production and release efficiency of minicell-encapsulated L-borneol.

In the first model, we used enzyme kinetics to understand the relationship between enzyme activity and substrate concentration. This model provides a framework to quantify how changes in substrate concentration affect the rate of reactions. By modeling the pathways involved in L-borneol production, we can simulate the production process and find our final yield based on the initial substrate concentration. Understanding this relationship will allow us to modify our production process and fine-tune L-borneol yield for maximum repellency and better aromas.

For our second model, since we're using minicell encapsulation to release L-borneol, it's essential to consider the transmembrane diffusion rate to assess how encapsulation influences both L-borneol release and its effective repellency. Since transmembrane diffusion rate is affected by the internal concentration of L-borneol, our second model utilizes the final yield obtained from our first model. This thorough understanding of L-borneol’s transmembrane movement is key to optimizing its production for maximum effectiveness.

Production: Kinetics

Method

There are two main pathways—the MVA pathway and the MEP pathway—and three downstream steps in the L-borneol production process. The following figure shows the entire reaction pathway for the production of L-borneol.


Figure 1. Overall pathway for L-borneol production

Glucose is the starting substrate in our pathway. However, as the pathway progresses, glucose is converted into various intermediates through a series of enzyme-substrate reactions. Each step is catalyzed by a specific enzyme that transforms the substrate into a new product, which then serves as the substrate for the next reaction. Eventually, L-borneol is produced.

Since the rate-limiting steps occur at a much slower rate than the non-rate-limiting steps, we assumed that the rate of the non-rate-limiting steps are negligible. Thus, we only modeled the rate-limiting steps, which are indicated by the red (overexpressed) enzymes shown in Figure 1. By chaining together the rate-limiting steps in the L-borneol production pathway, we are able to accurately approximate the rate of production of L-borneol. Our simplified version of the pathway only contains the rate-limiting steps, as seen in Figure 2.


Figure 2. The simplified version of our production pathway with only the rate-limiting steps (Clendening et al., 2010; Movahedi et al., 2021)

We utilized two different types of Michaelis-Menten equations in our production model. The first shows the equation for systems that contain only one substrate while the second describes a bi-substrate system.

\[ v = \frac{V_{\text{max}} \cdot [S]}{[S] + K_m} \]


\[ v = \frac{V_{\text{max}} [S_1] [S_2]}{K_m^{S_1} K_m^{S_2} + K_m^{S_1} [S_2] + K_m^{S_2} [S_1] + [S_1] [S_2]} \]

Variables

\( v \): production rate at a given time (unit: \( \mu M/sec \))
\( K_m \): the Michaelis-Menten rate constant. It is the substrate concentration when the reaction rate is half the maximum velocity. (unit: \( \mu M \))

\( V_{\max} \): the fastest rate the reaction can reach before the substrate concentration no longer affects the velocity of the reaction (unit: \( \mu M/sec \)).

Assumptions

  1. Only the rate-limiting steps in the MEP pathway affect the overall rate of production.

    1. The MEP pathway occurs at a slower rate than the MVA pathway.
    2. Since the end products of both the MEP and MVA pathways are IPP and DMAPP and the enzyme IDI regulates their equilibrium, one pathway can be prioritized.
    3. More enzymes are overexpressed in the MEP pathway, further indicating that it may occur at a slower rate.

  2. IPP and DMAPP are treated as a single substrate

    1. The IDI enzyme regulates the equilibrium between IPP and DMAPP and maintains a constant concentration ratio between the two substrates.

  3. Absence of inhibition

    1. We assume that no inhibitors are present to prevent the binding of substrates to enzymes, allowing for more straightforward enzyme-substrate interactions.

  4. G3P does not contribute to the production of Pyruvate

    1. G3P only contributes to the formation of DXP.
    2. G3P and Pyruvate are independent, non-interactive substrates.
    3. As both G3P and Pyruvate act as substrates in the first rate-limiting reaction of the MEP pathway, conversion of G3P to Pyruvate would overcomplicate enzyme kinetics equation for the reaction G3P + Pyruvate → DXP.

Michaelis-Menten Reaction Chain

We used the following equation to model the changes in substrate concentration over time for each step of the production of L-borneol.

\[ \frac{d[\text{Product}]}{dt} = \text{rate of formation} - \text{rate of consumption} \]

By implementing the equation above to each rate-limiting step in our pathway, we derive the chain of Michaelis-Menten equations below, with L-borneol as the final product. The only exception is the rate of formation of DXP, which follows a bi-substrate Michaelis-Menten equation, as both G3P and Pyruvate serve as substrates in the reaction (Dalwadi et al., 2018).


\[ \frac{d[DXP]}{dt} = \frac{V_{1max} [G3P][Pyruvate]}{K_{1m} K_{2m} + K_{1m} [Pyruvate] + K_{2m} [G3P] + [G3P] [Pyruvate]} - \frac{V_{2max} [DXP]}{[DXP] + K_{3m}} \]


\[ \frac{d[IPP \& DMAPP]}{dt} = \frac{V_{2max} [DXP]}{[DXP] + K_{3m}} - \frac{V_{3max} [IPP \& DMAPP]}{[IPP \& DMAPP] + K_{4m}} \]


\[ \frac{d[GPP]}{dt} = \frac{V_{3max} [IPP \& DMAPP]}{[IPP \& DMAPP] + K_{4m}} - \frac{V_{4max} [GPP]}{[GPP] + K_{5m}} \]


\[ \frac{d[BPP]}{dt} = \frac{V_{4max} [GPP]}{[GPP] + K_{5m}} - \frac{V_{5max} [BPP]}{[BPP] + K_{6m}} \]


\[ \frac{d[L\text{-borneol}]}{dt} = \frac{V_{5max} [BPP]}{[BPP] + K_{6m}} \]

Obtaining Constants

Obtaining the constants in our Michaelis-Menten equations, \( K_m \) and \( V_{\max} \), for each rate-limiting step proved to be a significant bottleneck to this model. While we were able to find most of the \( k_{\text{cat}} \) and \( K_m \) values for the MVA and MEP pathways from literature or enzyme databases such as BRENDA, there existed discrepancies between available data and our desired experimental conditions. These conditions include the environment in which the experiment was conducted (such as temperature and pH), the organism used, and the specific substrate involved in the reaction. In the calculations that we have done, we only utilized the constants that we believe are accurate for our model.

In addition to the kinetic constants, the value of \( V_{\max} \) is essential for our equations, and it can be calculated with the equation:

\[ V_{\max} = k_{\text{cat}} \cdot [E_t] \]

Since we obtained the values of \( k_{\text{cat}} \) through research, the only missing variables are the total concentrations of the rate-limiting enzymes.

We determined the total enzyme concentration by running a 10% SDS page for all the enzymes utilized in the system. Using a computer software called ImageJ, we measured the relative luminescence of each rate-limiting enzyme in comparison to the total luminescence for the lane. As a reference, we added a known quantity of Bovine Serum Albumin (BSA), 1 µg in 5 µL, to each enzyme sample. Below is an image of the SDS page.


Figure 3. Annotations of the enzymes on the 10% SDS page

We found that BSA has a concentration of 3.01 µM. To find the concentration for the other enzymes, we normalized the ratio for BSA and the other enzymes as we already know the concentration for BSA. This process was repeated for two induced trials, which were then averaged to get the final concentration to ensure higher accuracy. This is the calculation performed to obtain the concentration of BSA:

\[ 1 \times 10^{-6} \, \text{g} \times \frac{1 \, \text{mol}}{66500 \, \text{g}} \times \frac{1}{5 \times 10^{-6} \, \text{L}} = 3.01 \times 10^{-6} \, \text{M} = 3.01 \, \mu \text{M of BSA} \]

We obtained the relative ratio for each enzyme in comparison to BSA by getting the luminescence of the enzyme and dividing it by the luminescence of BSA. We then used the ratio and multiplied it by the known concentration for BSA. The following is the calculation for HmgR:

\[ \frac{1202287.166}{1049993.663} \times 3.01 \, \mu M = 3.45 \, \mu M [\text{DXS}] \]

The following list shows the total enzyme concentration for each rate-limiting enzyme:


Table 1. The corresponding \( K_m \), \( k_{\text{cat}} \), \( V_{\max} \), and concentration values for each rate-limiting reaction (Koppisch et al., 2002)

In addition, we can use the \( K_m \) and \( V_{\max} \) values to evaluate the enzyme's efficiency in catalyzing the reaction. Catalytic efficiency is calculated through \(\frac{k_{\text{cat}}}{K_m}\).

  • When \( K_m \) is large, the enzyme has a low affinity and therefore can only reach \( V_{\max} \) at a high substrate concentration, making the reaction less efficient.
  • When \( K_m \) is low, the enzyme has a high affinity and therefore can reach \( V_{\max} \) at a lower substrate concentration, making the reaction more efficient.
  • When \( k_{\text{cat}} \) is large, the catalytic efficiency is high because the rate for product formation, given by \( \frac{d[P]}{dt} = k_{\text{cat}} [\text{ES}] \), is higher. This indicates that the enzyme is more efficient in catalyzing the reaction to form the product.
  • When \( k_{\text{cat}} \) is small, the catalytic efficiency is low because the rate for product formation, given by \( \frac{d[P]}{dt} = k_{\text{cat}} [\text{ES}] \), is lower. This indicates that the enzyme is less efficient in catalyzing the reaction to form the product.

To summarize, if \( K_m \) is high or \( k_{\text{cat}} \) is low, the reaction is less efficient while if \( K_m \) is low or \( k_{\text{cat}} \) is high, the reaction is more efficient. This supports our assumption that the rate-limiting steps in the MEP pathway more significantly influence the overall production rate of L-borneol. The higher \( K_m \) value for the rate-limiting step in the MEP pathway (90 to 140 µM) compared to the MVA pathway (20 µM) further indicates that the MEP pathway is less efficient.

Model Simulation

After obtaining our constants, we solved our set of differential equations using Python, which allowed us to calculate the overall production rate of L-borneol and generate a graph showing the concentration changes of substrates over time at different steps of the pathway. However, it is important to note that several constants inputted in our code are placeholders. Despite not having specific kinetic parameters, the graph generated by our model simulation helps us to understand the relationship between the initial glucose concentration and the final yield of L-borneol. This allows us to determine the optimal initial glucose concentration needed to achieve our desired amount of L-borneol.

Our Python code uses the ODEINT function to solve the system of ordinary differential equations (ODEs). The reaction chain models how each variable interacts with one another in the overall system. The solver returns the substrate concentrations at specific time points, then, integrates these values into the next equation at subsequent time points (100 time points within 10-second intervals), capturing how each substrate concentration changes over time.


def michaelis_menten(S, Vmax, Km):
    return Vmax * S / (Km + S)


def michaelis_menten_2vars (Sa, Sb, V2max, Kam, Kbm):
    return V2max*Sa*Sb/(Kam*Kbm +Kam*Sa + Kbm*Sa + Sa*Sb)


def reaction_chain(y, t, V1max, K1m, K2m, V2max, K3m, V3max, K4m, V4max, K5m, V5max, K6m):
    G3P, Pyruvate, DXP, IPPandDMAPP, GPP, BPP, Lborneol = y

    # Rate of G3P +Pyruvate -> DXP
    rateA = michaelis_menten_2vars(G3P, Pyruvate, V1max, K1m, K2m)

    # Rate of DXP -> IPP and DMAPP
    rate1 = michaelis_menten(DXP, V2max, K3m)


    # Rate of IPP and DMAPP -> GPP
    rate2 = michaelis_menten(IPPandDMAPP, V3max, K4m)

    # Rate of GPP -> BPP
    rate3 = michaelis_menten(GPP, V4max, K5m)


    # Rate of BPP -> L-borneol
    rate4 = michaelis_menten(BPP, V5max, K6m)


    # ODEs
    dG3P_dt = - 1*rateA
    dPyruvate_dt = - 1*rateA
    dDXP_dt = rateA - rate1 
    dIPPandDMAPP_dt = rate1 - rate2 
    dGPP_dt = rate2 - rate3  
    dBPP_dt = rate3 - rate4
    dLborneol_dt = rate4


# Time points (0 to 10 seconds, 100 steps)
t = np.linspace(0, 10, 100)


# Solve ODEs, input with constants e.g. Vmax & Km
solution = odeint(reaction_chain, initial_concentrations, t, args=(V1max, K1m, K2m, V2max, K3m, V3max, K4m, V4max, K5m, V5max, K6m))

                            

Figure 4. This graph depicts the concentration of rate-limiting substrates over time(constants found using XJTU-China's ProtGPT-2 model)

Limitations

There are multiple limitations that we were unable to model:

  1. Neglecting inhibition when the enzyme-substrate complex is bounded by an inhibitor

    1. Inhibition can affect the rate of production by temporarily or permanently reducing enzyme activity and halting the reaction. To obtain more accurate results, an extended Michaelis-Menten model that incorporates the effects of inhibitors is needed.

  2. Inability to determine multiple kinetic parameters

    1. Insufficient research was done on our engineered downstream steps, so there are limited available literature data on the values of constants.
      1. We utilized the BRENDA database; however, each recorded constant is obtained under specific conditions that do not perfectly align with our experiments.
    2. Due to insufficient technology, we were unable to determine the kinetic constants from experimental data.
    3. Further exploration of constants through experimental data would allow us to more accurately model the rate of production of L-borneol.

Future Experiments

Since we were unable to find the necessary constants for modeling the production of L-borneol from literature, our future plan is to obtain these constants using experimental data. There are two regression models we could use to find our constants: the Lineweaver Burk plot and the Eadie Hofstee plot. Both plots require experimental data of different substrate concentrations and their corresponding initial reaction rates (Blaber, 2021).

Lineweaver Burk Plot

\[ \begin{aligned} & \frac{1}{v}=\frac{K_m+[S]}{V_{\max }[S]} \\ & \frac{1}{v}=\left(\frac{K_m}{V_{\max }}\right)\left(\frac{1}{[S]}\right)+\frac{1}{V_{\max }} \\ & y=m x+b \end{aligned} \]

Lineweaver Burk Plot

\[ \begin{aligned} & y = \frac{1}{v} \\ & m = \frac{K_m}{V_{\max }} \\ & x = \frac{1}{[S]} \\ & b = \frac{1}{V_{\max }} \\ & \text{y-intercept: } \frac{1}{V_{\max }} \\ & \text{x-intercept: } -\frac{1}{K_m} \\ & \text{slope: } \frac{K_m}{V_{\max }} \end{aligned} \]

Eadie-Hofstee Plot

\[ \begin{aligned} & y = m x + b \\ & v = -K_m\left(\frac{v}{[S]}\right) + V_{\max} \end{aligned} \]

\[ \begin{aligned} & y = v \\ & m = -K_m \\ & x = \frac{v}{[S]} \\ & b = V_{\max } \\ & \text{y-intercept: } v_{\max } \\ & \text{x-intercept: } \frac{V_{\text{max}}}{K_m} \\ & \text{slope: } -K_m \end{aligned} \]


























The Lineweaver Burk Plot utilizes the reciprocal values of both substrate concentration and the initial reaction rates to create a linear \(\frac{1}{[S]}\) vs. \(\frac{1}{V_{\max }}\) graph. This graph allows the constants \( V_{\max} \) and \( K_m \) to be easily interpreted through computer-generated regression.

To find \( V_{\max} \), we can take the reciprocal of the y-intercept \(\frac{1}{V_{\max }}\). To find \( K_m \), we can substitute the previously calculated \(\frac{K_m}{V_{\max }}\) value into the slope of the graph, \(\frac{K_m}{V_{\max }}\), and solve for \( K_m \). The calculations for these two parameters are demonstrated below:

\[ \begin{aligned} & V_{\max} = \frac{1}{b} \\ & K_m = m \cdot V_{\max} \end{aligned} \]

In the Eadie-Hofstee Plot, the x-axis represents the ratio of velocity to substrate concentration (\(\frac{v}{[S]}\)) while the y-axis represents the initial velocity \(v\).This graph also allows the constants \( V_{\max} \) and \( K_m \) to be easily interpreted through computer-generated regression.

To find \( V_{\max} \), we can locate the y-intercept of the graph. To find \( K_m \), we can take the absolute value of the slope.

By applying either the Lineweaver Burk Plot or the Eadie-Hofstee Plot to the rate-limiting steps of our L-borneol production pathway, we can experimentally obtain the necessary kinetic constants of \( V_{\max} \) and \( K_m \). Then, these values will be utilized in our chain of Michaelis-Menten equations (Tiwardi, n.d.).

Release: Membrane Diffusion

Through our enzyme kinetics model, we can theoretically determine the final yield of L-borneol, which in turn allows us to know its internal concentration within our minicells. This information is critical for accurately modeling the release of L-borneol from minicells.

The transmembrane diffusion rate depends on factors like distance, diffusant size, and velocity. A key factor is the concentration gradient: a steeper gradient drives faster diffusion, but as the internal concentration decreases, the diffusion rate declines (“Factors affecting the rate of diffusion,” 2024). In typical models, internal and external concentrations are constant. However, in our case, borneol production within minicells adds a second variable that impacts internal concentration and thus the diffusion rate. To accurately capture this dynamic, our model incorporates the ongoing borneol production's effect on the concentration gradient.

Method

The parameters of our model include:

\[ \begin{aligned} & C(t): \text{ Concentration of borneol inside the minicell at time } t \\ & R: \text{ Rate of internal production of borneol} \\ & D: \text{ Diffusion coefficient of borneol} \\ & V: \text{ Volume of the minicell} \\ & A: \text{ Surface area of the minicell} \\ & L: \text{ Thickness of the minicell membrane} \end{aligned} \]

An elementary way of modeling the concentration gradient of borneol across the minicell membrane would be \( \frac{dC}{dx} = \frac{C_{\text{inside}} - C_{\text{outside}}}{L} \) , the difference of the internal and external borneol concentration divided by the membrane thickness, \(L\).


In this scenario, due to the lack of borneol in the environment, we assume \(C_{\text{outside}}\) to be close to zero. With this assumption, our former equation can be modified into: \[ \frac{dC}{dx} \approx \frac{C_{\text{inside}}}{L} \]


Following this, \( \frac{C_{\text{inside}}}{L} \) can be substituted into Fick’s First Law, \(J = -D \frac{dC}{dx}\), to obtain: \[ J = -\frac{D C_{\text{inside}}}{L} \]


The equation can further be modified by adjusting the diffusion flux from the amount of borneol diffused per area per unit time to the amount of substance per area per unit time at time \(t\): \[ J(t) = -\frac{D}{L} C(t) \]


This is then multiplied by \(A\), the area of diffusion, and transformed into \( J(t) \cdot A = -\frac{AD}{L} C(t), \) the rate of the outward diffusion of borneol at time \(t\).


On a closely related subject, another vital model concept is as follows: \[ \text{change in concentration at time } t = \frac{\text{rate of production at time } t + \text{rate of exit at time } t}{\text{minicell volume}} \]


To find the change in the internal concentration of borneol at time \(t\), we need to combine the negative rate at which borneol diffuses out of the minicell and the rate at which borneol is produced within the cell at time \(t\) and divide the sum by minicell volume \(V\).


The rate of borneol production, \(R\), is assumed to be constant throughout the entire lifespan of the minicell. After substituting all of the variables, the following equation is produced: \[ \frac{dC(t)}{dt} = \frac{R - A \cdot D \cdot \frac{C(t)}{L}}{V} \]


The equation above can then be rearranged and simplified to solve for \(C(t)\), the internal concentration of borneol at time \(t\): \[ C(t) = \frac{R \cdot L}{AD} + C_0 \cdot e^{-\frac{AD}{LV} t} \]


The constant of integration, \(C_0\), is assumed to be the initial condition of the minicell at time \(t = 0\). Inputting \(t = 0\) into our most recent equation generates the equation below: \[ C(0) = \frac{R \cdot L}{AD} + C_0 \cdot e^{-\frac{AD}{LV} \cdot 0} = \frac{R \cdot L}{AD} + C_0 \]


After a brief simplification, we get: \[ C(0) = \frac{R \cdot L}{AD} + C_0 \]

This can then be reordered to produce: \[ C_0 = C(0) - \frac{R \cdot L}{AD} \]


And we can substitute \(C_0\) into our former solved function \(C(0)\) and replace the constants \(\frac{AD}{L}\) with \(K\) for the following equations: \[ C(t) = \frac{RL}{AD} + \left(C(0) - \frac{RL}{AD}\right) e^{-\frac{AD}{LV}t} \] \[ C(t) = \frac{R}{K} + (C(0) - \frac{R}{K}) e^{-\frac{K}{V} t} \]


The original equation for the rate of exit for borneol, \(J(t) \cdot A = \frac{AD}{L} C(t) = K \cdot C(t)\), can be updated by substituting \(C(t)\) with our solved equation. We get: \[ \text{rate of exit at time } t = K \cdot \left( \frac{R}{K} + (C(0) - \frac{R}{K}) e^{-\frac{K}{V} t} \right) \]


We then take the integral of this equation, \(\int_0^t K \cdot C(t) \, dt\), to obtain the amount of borneol that diffused out of the minicell after time \(t\) passed: \[ \int_0^t K \cdot C(t) \, dt = R \cdot t - V(C(0) - \frac{R}{K}) e^{-\frac{K}{V} t} + c \]


The constant of integration, \(C\), as we mentioned beforehand, is the initial concentration of borneol outside the minicell. Thus, we have the final equation, which is as follows: \[ \int_0^t K \cdot C(t) \, dt = R \cdot t - V(C(0) - \frac{R}{K}) e^{-\frac{K}{V} t} + C(0) \]

Obtaining Parameters


The diffusion coefficient of borneol is a variable that is widely used within our equations. We determined it through the Stokes-Einstein-Sutherland equation, which is applied for spherical diffusants within liquids (Baer et al., 2023). In our experiment, we assume the geometry of borneol to be spherical, which makes modeling it easier and thus giving us the equation for its diffusion coefficient below, where: \[ D = \frac{k_B T}{6 \pi \mu R} \] with the variables:

  • \( k_B \): Boltzmann’s constant
  • \( T \): temperature (K)
  • \( \mu \): solvent viscosity
  • \( R \): solute radius

  • In our model, the temperature will be set at room temperature, averaging around 23℃ or 296.15K, as this will be the average temperature in which our product is used (“The climate of Taiwan,” n.d.).

  • The solvent in our scenario would be the cell membrane, as it is the medium in which our solute, borneol, diffuses through. The viscosity of our minicell membrane was determined to be around 950 centipoise (Mika et al., 2016).

  • The estimation of borneol’s radius was done through the density equation, \( d = \frac{m}{V_b} = \frac{M_r}{n \cdot V_b} \), where the density of borneol is equal to its mass over volume, or more specifically, its molecular mass over its volume multiplied by Avogadro’s number.

    Rewriting the equation gives us \( V_b = \frac{M_r}{n \cdot d} \), and \( V_b \), the volume of borneol, can be substituted with the standard equation for calculating the volume of spheres. Again, we make the simplification that borneol’s geometry is spherical. We thus get the equation below: \[ \frac{4}{3} \pi r^3 = \frac{M_r}{n \cdot d} \]


    Reordering the equation to solve for \( r \) then gives us our final equation: \[ r = \sqrt[3]{\frac{3 M_r}{4 \pi n \cdot d}} \]

    We can substitute the values for the density and molar mass of borneol, both of which have been determined through external research (PubChem, n.d.), to get: \[ r = \sqrt[3]{\frac{3 \times (154.25)}{4 \pi \times (6.022 \times 10^{23}) \times 1.011}} = 3.9 \times 10^{-8} \, \text{cm/molecule} = 0.39 \, \text{nm/molecule} \]

    Thus, the radius of our solute, borneol, was approximated to be 0.39 nm.


Table 2. Summary of values needed to find borneol’s diffusion coefficient

Thus with all variables obtained, we can solve for the diffusion coefficient of borneol. Solving for it with the Stokes–Einstein–Sutherland equation, we get \( D = 0.005894 \times 10^{-10} \, \text{m}^2/\text{s} \).


Another parameter needed would be \( V \), the volume of our minicells. As most research describes the size of minicells by their width (diameter) and not their volume, we would have to calculate the value ourselves with the equation: \[ V = \frac{4}{3} \pi r^3 = \frac{4}{3} \pi \left(\frac{d}{2}\right)^3 \]


Minicell sizes can vary significantly in different scenarios; their diameters have been recorded to range anywhere from 100–400 or 800 nm (Kim et al., 2022; Yu et al., 2021). We chose to simulate the volume of our minicells with the assumed diameter of 800 nm, as proving the efficacy of our product with the most extreme scenario (having the largest solute and therefore the slowest diffusion speed) will mean that our product can function in all other situations.

Inputting \( d = 800 \, \text{nm} \) (or \( r = 400 \, \text{nm} \)) into the equation gives us the approximation for the maximum volume of our minicell: \[ V = \frac{4}{3} \pi \cdot 400^3 \] \[ V \approx 0.268 \, \mu m^3 = 2.68 \times 10^{-19} \, m^3 \]


Additionally, the radius of our minicell can be employed to also calculate each minicell’s surface area. Using the equation \( A = 4 \pi r^2 \), the surface area of an individual minicell is approximately: \[ A = 4 \pi \cdot 400^2 \] \[ A = 2.01 \, \mu m^2 = 2.01 \times 10^{-12} \, m^2 \]


Lastly, the thickness of our minicell membranes, \( L \), is determined to be 7.5 nanometers, or \( 7.5 \times 10^{-9} \, \text{meters} \) (“Thickness of cytoplasmic and outer membrane,” n.d.).


Ultimately, the constant \( K \), which is equal to \( \frac{AD}{L} \), will be \( K = 1.58 \times 10^{-16} \).


Thus, with all parameters obtained, they can be substituted into our model to produce the final version: \[ R \cdot t - \left(2.68 \times 10^{-19}\right)(C(0) - \frac{R}{1.58 \times 10^{-16}}) e^{-589.55t} \]

or

\[ R \cdot t - V(C(0) - \frac{R}{K}) e^{-\frac{K}{V} t} + C(0) \]

where:

  • \( R \): Rate of internal production of L-borneol
  • \( V \): Volume of the minicell
  • \( K \): \( \frac{A \cdot D}{L} \), diffusion constant
  • \( D \): Diffusion coefficient of L-borneol
  • \( A \): Surface area of the minicell
  • \( L \): Thickness of the minicell membrane
  • \( C(0) \): Initial internal concentration of L-borneol


Model Simulation

After writing a Python code, we were able to plot the external concentration of L-borneol over time given any initial internal concentration and internal production rate.


Figure 5. The external concentration of borneol over time (placeholder value for initial concentration)

Results & Conclusion

The lack of key kinetic constants from the production model has prevented us from accurately determining the L-borneol production rate. This has also affected the diffusion model from connecting with the production model. In the absence of an accurate L-borneol production rate, the diffusion dynamics of L-borneol cannot be precisely simulated. However, our model demonstrates the relationship between the initial glucose concentration inputted and the amount of L-borneol diffused if further constants are obtained.

Limitations

It is known that the rate of diffusion is affected by L-borneol concentration. As the initial concentration of L-borneol increases, the diffusion coefficient also increases at an exponential rate. This is unintuitive as the coefficient is normally a constant that does not change with concentration; however, L-borneol’s ability to disrupt the lipid bilayers makes it so that increased concentrations increase the diffusion coefficient, and, therefore, the rate of membrane diffusion. Nonetheless, we can make the inference that our product would only be even more effective when the rate of diffusion increases together with L-borneol concentration.

  1. Baer, A., Wawra, S. E., Bielmeier, K., Uttinger, M. J., Smith, D. M., Peukert, W., Walter, J., & Smith, A. (2023). The Stokes–Einstein–Sutherland equation at the Nanoscale revisited. Small, 20(6). https://doi.org/10.1002/smll.202304670
  2. Blaber, M. (2021, August 23). 7.2: Derivation of Michaelis-Menten equation. Biology LibreTexts. https://bio.libretexts.org
  3. Clendening, J. W., Pandyra, A., Boutros, P. C., El Ghamrasni, S., Khosravi, F., Trentin, G. A., Martirosyan, A., Hakem, A., Hakem, R., Jurisica, I., & Penn, L. Z. (2010). Dysregulation of the mevalonate pathway promotes transformation. Proceedings of the National Academy of Sciences of the United States of America, 107(34), 15051–15056. https://doi.org/10.1073/pnas.0910258107
  4. Dalwadi, M. P., Garavaglia, M., Webb, J. P., King, J. R., & Minton, N. P. (2018). Applying asymptotic methods to synthetic biology: Modeling the reaction kinetics of the mevalonate pathway. Journal of Theoretical Biology, 439, 39–49. https://doi.org/10.1016/j.jtbi.2017.11.022
  5. Factors affecting the rate of diffusion - Transport into and out of cells - AQA Synergy - GCSE Combined Science Revision - AQA Synergy - BBC Bitesize. (2024, June 6). BBC Bitesize. https://www.bbc.co.uk/bitesize/guides/ztrgng8/revision/2
  6. Kim, S., Chang, W., & Oh, M. (2022). Escherichia coli minicells with targeted enzymes as bioreactors for producing toxic compounds. Metabolic Engineering, 73, 214–224. https://doi.org/10.1016/j.ymben.2022.08.006
  7. Koppisch, A. T., Fox, D. T., Blagg, B. S., & Poulter, C. D. (2002).E. coliMEP synthase: steady-state kinetic analysis and substrate binding. Biochemistry, 41(1), 236–243. https://doi.org/10.1021/bi0118207
  8. Mika, J. T., Thompson, A. J., Dent, M. R., Brooks, N. J., Michiels, J., Hofkens, J., & Kuimova, M. K. (2016). Measuring the viscosity of the Escherichia coli plasma membrane using molecular rotors. Biophysical Journal, 111(7), 1528–1540. https://doi.org/10.1016/j.bpj.2016.08.020
  9. Movahedi, A., Wei, H., Pucker, B., Sun, W., Li, D., Yang, L., & Zhuge, Q. (2021, January 1). Regulation of poplar isoprenoid biosynthesis by methylerythritol phosphate and mevalonic acid pathways interactions. bioRxiv. https://www.biorxiv.org/content/10.1101/2020.07.22.216804v3.full
  10. PubChem. (n.d.). Borneol, (-)-. PubChem. https://pubchem.ncbi.nlm.nih.gov/compound/1201518#section=Depositor-Supplied-Synonyms
  11. The climate of Taiwan. (n.d.). http://twgeog.ntnugeog.org/en/climatology/
  12. Thickness of cytoplasmic and outer membrane - Bacteria Escherichia coli - BNID 109060. (n.d.). https://bionumbers.hms.harvard.edu/bionumber.aspx?id=109060&ver=0&trm=e+coli+membrane+thickness&org=
  13. Tiwari, V. (n.d.). Molecular Enzymology and Protein Engineering: Determination of kinetics parameters using Lineweaver-Burk, Eadie-Hofstee plot, Scatchard plot. e-PG Pathshala for Biophysics, MHRD project, UGC. https://epgp.inflibnet.ac.in/epgpdata/uploads/epgp_content/S001174BS/P001200/M010877.pdf
  14. Tiwari, V. (n.d.). Molecular Enzymology and Protein Engineering: Determination of kinetics parameters using Lineweaver-Burk, Eadie-Hofstee plot, Scatchard plot. e-PG Pathshala for Biophysics, MHRD project, UGC. https://epgp.inflibnet.ac.in/epgpdata/uploads/epgp_content/S001174BSTextNov27.pdf
  15. Yu, H., Khokhlatchev, A. V., Chew, C., Illendula, A., Conaway, M., Dryden, K., Maeda, D. L. N. F., Rajasekaran, V., Kester, M., & Zeichner, S. L. (2021). Minicells from highly genome reduced Escherichia coli: Cytoplasmic and surface expression of recombinant proteins and incorporation in the minicells. ACS Synthetic Biology, 10(10), 2465–2477. https://doi.org/10.1021/acssynbio.1c00375



Stickiness Metrics Framework

Overview

The stickiness of traditional oil-based mosquito repellents is a massive deterrent to people who are bothered by their discomfort. According to our data, more than 75% of people surveyed indicated that the stickiness of traditional mosquito repellents discouraged them from using them. This is a significant portion of the population that we would like to appeal to with BOROHMA.

Our Stickiness Metrics Framework aims to prove that our product is more comfortable than traditional oil-based repellents by accurately quantifying the comfort level of our product. We decided to approach the issue from two different angles: viscosity and objective stickiness. In the viscosity section, we constructed a time-based viscosity model for colloidal solutions to obtain a theoretical evaluation of our product’s comfortability. In the objective stickiness section, we combined the previously made viscosity equation with a stickiness equation that takes into account the effects of adhesion, cohesion, and particle velocity to accurately model the stickiness of colloid solutions.

Figure 1. Survey on how the sticky feel of mosquito repellents affect its usage

Time-Based Viscosity

We begin our model by modeling viscosity. While adhesion and cohesion are the major determining factors for stickiness, viscosity is the major determining factor for adhesion and cohesion.

Methods

Our first step towards accurately quantifying stickiness involves measuring the viscosity of our minicell solution over time. As the water from our solution evaporates, the ratio of our minicells to the solution increases, thus increasing the viscosity of the repellent. We model this increasing viscosity, which allows us to determine if our repellent becomes uncomfortably viscous as time passes. We decided to utilize the flow behavior index for the viscosity of suspensions because it can accurately represent the viscosity of the solution even at extremely high concentrations ("The Influence of Particles on Suspension Rheology," n.d.).

The flow behaviour index is derived based on empirical observations of how viscosity changes with the concentration of particles in a colloidal suspension. Viscosity (\(\eta\)) is defined as a measure of a fluid's resistance to flow. In colloidal suspensions, the viscosity is influenced by the volume fraction of solid particles (\(\phi\)) in the fluid. As the concentration of particles increases, the interactions between them become significant, affecting the overall viscosity. The term \(\eta_0\) ​represents the viscosity of the fluid for the initial volume fraction. This serves as the baseline viscosity from which changes due to particle concentration are measured(Xia & Krueger, 2022).

The general form of the relationship is often expressed as:

\[ \eta = \eta_0 \cdot (1 + f(\phi)) \]

where \(f(\phi)\) is a function that describes how viscosity increases with concentration.

We can use the power law form for the function \(f(\phi)\) which is a common choice which captures the non-linear increase in viscosity with concentration:

\[ f(\phi) = k \cdot \phi(t)^n \]

Here, k is a constant that reflects the sensitivity of viscosity to changes in particle concentration, and n is an exponent that characterizes the nature of the relationship. This leads to the equation:

\[ \eta(t) = \eta_0 \cdot \left(1 + k \cdot \phi(t)^n\right) \]

We can combine this with the volume fraction of minicells at time t which we derive from the total volume at time t:

The total volume at time t is:

\[ V(t) = V_{w0} + V_c - E \cdot t \]

Where the total volume of the solution is the volume of the minicells plus the volume of the initial water minus the evaporation rate times time.

We can combine the total volume with the volume fraction of minicells to get the volume fraction of minicells at time t:

\[ \phi(t) = \frac{V_c}{V(t)} = \frac{V_c}{(V_{w0} + V_c - E \cdot t)} \]

Combining this into the previously derived function n(t) we get:

\[ \eta(t) = \eta_0 \cdot \left(1 + k \cdot \left(\frac{V_c}{(V_{w0} + V_c - E \cdot t)}\right)^n\right) \]

  • \(V_0\): Initial volume of the water (ml)
  • \(V_c\): Volume of the minicells (ml)
  • \(E\): Evaporation rate (volume of water evaporating per minute in ml)
  • \(\eta_0\): initial viscosity of the solution (in cP)
  • \(t\): Time (in minutes)
  • \(k\): Reflects how much the viscosity increases per unit increase in the volume fraction
  • \(n\): power-law exponent that determines the nature of the relationship between concentration and viscosity

Experimental Assumptions

We made various assumptions to simplify the modeling process while keeping the assumptions logical to minimize discrepancies. One assumption we made is that the market repellent we used to compare our repellent to is comparative in stickiness to other market repellents that are oil-based. We also assumed that the main contributing factor to the stickiness of our repellent were the minicells in our repellent. This is a logical assumption as smaller molecules such as borneol and fragrance particles are unlikely to affect viscosity.

Viscosity Experiments

To find constants \(\eta\), \(k\), and \(\eta_0\) using experimental data, we will use the viscosity test called the inclined plane method. This test involves finding the time it takes for our minicell solution to flow down a set slope and comparing the ratio of these times to water which has a viscosity of 1cP. Using these methods we will determine the viscosity of our minicell solution and another DEET-based repellent using pure water as a control. We will test our minicell solution's viscosity at 10%, 15%, 20%, 25%, and 30% concentrations and regress the outputs on an exponential graph.

To prove that we can use the inclined plane method to accurately model the viscosity we show that for a fluid moving down an inclined plane, the motion is driven by gravity and opposed by viscous forces. Assuming our fragrance and market repellent behave as a normal Newtonian fluid, we can derive a relationship between flow velocity and viscosity (Sewalt et al., 2020).

The key forces involved are:


  • Gravitational Force Component (Fgravity):The component of gravitational force acting along the incline, which accelerates the fluid

\[ F(\text{gravity}) = \rho V \operatorname{g \sin}(\theta) \]

Where:

  • ρ is the fluid density
  • V is the volume of the fluid
  • G is the gravitational acceleration
  • θ is the angle of the incline

  • Viscous Drag Force (Fviscosity):The viscous resistance opposing the flow, which is proportional to the fluid's viscosity and the velocity gradient (shear rate) in the fluid:

\[ F(\text{gravity}) = \eta \frac{\partial u}{\partial y} \]

Where:

  • η is the dynamic viscosity of the fluid
  • \(\frac{\partial u}{\partial y}\)​ is the velocity gradient across the flow

In the case of a simple fluid drop sliding down the inclined plane, the velocity of the fluid will reach a steady state where the gravitational force pulling the fluid down the plane is balanced by the viscous drag resisting the motion.

The time (T) it takes for the fluid to move a certain distance (L) down the incline is inversely proportional to the flow velocity (u):

\[ T \propto \frac{L}{u} \]

Since the velocity of the fluid is a function of the ratio between gravitational force and viscous drag, we can express the velocity as:

\[ u \propto \frac{ρgsin⁡(θ)}{\eta} \]

Thus, the time T is proportional to the viscosity:

\[ T \propto \frac{\eta}{ρgsin⁡(θ)} \]

For two different fluids flowing under the same conditions, we can compare their viscosities directly by comparing their flow times. The other variables (ρ, g, sin⁡(θ)) cancel out when taking the ratio of the times for the two fluids.

Thus we can use this equation to find the viscosity of our repellent fragrances at different concentrations:

\[ \frac{\text{Travel Time of target solution }(s)}{\text{Travel Time of water }(s)} \cdot \text{Base viscosity of water }(1 \, cP) = \text{Viscosity of target solution }(cP) \]

Figure 2. Demonstration of our lab procedure when testing viscosity

Table 1. Results of our experimental viscosity tests of our repellent at different concentrations as well as water and market repellent solution

Figure 3. Exponential regression of our experimental data points using Desmos

Figure 4. Cubic and exponential regression using Python to find constants \(k\) and \(n\)

We constructed a code that can regress our equation \[ \eta(t) = \eta_0 \cdot \left(1 + k \cdot \left(\frac{V_c}{(V_{w0} + V_c - E \cdot t)}\right)^n\right) \] onto our data points to find the values K and N. We did this by simplifying our power-law equation to \[ \eta(t) = \eta_0 \cdot \left(1 + k \cdot \phi(t)^n\right) \] where \(\phi\) is the volume fraction which acts as our independent variable from the cubic regression of our experimental data. We can do this because \(\phi\) is equivalent to (\(\frac{V_c}{V_{w0} + V_c - E \cdot t}\)) at a certain point in time t.


def cubic_model(phi, a, b, c, d):
    return a * phi**3 + b * phi**2 + c * phi + d
initial_guess_cubic = [1, 1, 1, 1]
popt_cubic, _ = curve_fit(cubic_model, phi_data, eta_data, 
p0=initial_guess_cubic)        
                            

We use the cubic function because the R squared value (0.99>0.94)is higher than the exponential function. Thus we can use the synthetic data from our cubic model to model a finer range of \(\phi\) values and plug it into our power law.


initial_guess_power = [0.1, 1.0]  # Initial guesses for k and n
popt_power, _ = curve_fit(power_law_model, phi_fit, eta_fit_cubic, p0=initial_guess_power)
k, n = popt_power
print(f"Fitted power-law parameters: k = {k:.4f}, n = {n:.4f}")       
                            

To properly regress the values of \(k\) and \(n\) we start off with the values 0.1 and 1.0 respectively so that as their values increase the exponential function is slowly fitted more closely to our actual data points. This allows us to accurately print the final values of k and n of our exponential model that is fitted to the experimental data.

Results

Using the \(k\) and \(n\) constants derived from our experimental data and plugging in our other values we can obtain the final equation:

\[ \eta(t) = 1.2 \cdot \left(1 + 3.81 \left(\frac{30}{150 - t}\right)^{1.78}\right) \]

  • \(V_0\): Initial volume of the water which is (1ml per spray)*(5 sprays)*(80% of the solution)=4ml
  • \(V_c\): Volume of the minicells which is (1ml per spray)*(5 sprays)*(20% of the solution)=1ml
  • \(E\): Evaporation Rate (volume of water evaporating per unit time assume 0.5ml/min)
  • \(\eta_0\): initial viscosity of the solution(starting viscosity of our repellent which is 1.2 cP)
  • \(t\): Time (in minutes)
  • \(k\) = 3.81: reflects how much the viscosity increases per unit increase in the volume fraction
  • \(n\) = 1.78: power-law exponent that determines the nature of the relationship between concentration and viscosity

Thus inputting our values into our equation we get a viscosity over time graph that shows us how the viscosity changes as the water evaporates when our repellent is sprayed on the skin.

Figure 5. Graph showing when our repellent fragrance reaches 2.026 cP using python

Figure 5 shows that our repellent's viscosity when sprayed on skin only reaches 2.026 cP after more than 5 minutes assuming the average evaporation rate of water is 0.5ml/min. This proves that our repellent is significantly less viscous than an oil-based market repellent even as the water evaporates and volume fraction increases.

Figure 6. Comparison of the viscosity over time graphs of our repellent fragrance vs. a market repellent

If we were to assume that the market repellent’s viscosity per volume fraction increases at the same rate as our repellent except with a slower rate of evaporation (oil evaporates slower than water) our repellent would only become as viscous as the market repellent after more than 7 minutes and the final viscosity would be more than 4cP lower. Since viscosity is directly correlated to stickiness as proved by the second part of our model this indirectly proves that our repellent is more comfortable than oil-based market repellents.


def modified_viscosity_over_time(t, n0, k, Vc, Vw0, E, n):
    total_water = Vw0 + Vc
    remaining_water = np.maximum(total_water - E * t, 0)  
    phi = np.where(remaining_water > 0, Vc / remaining_water, 1)  
    baseline = k * (Vc / total_water) ** n
    viscosity = n0 * (1 + k * phi ** n - baseline)
    return np.minimum(viscosity, viscosity[-1])      
                            

Since our repellent will have an average minicell concentration of about 20 percent we want to set our initial volume fraction to 1.2. To fit our equation better we will subtract our baseline value so that the graph will always start at \(n_{\text{0}}\). We also used np.maximum(total_water - E * t, 0) and np.minimum(viscosity, viscosity[-1]) to plateau the viscosity at its maximum value and ensure the remaining water doesn't go negative.

Objective Stickiness Equation

After determining the viscosity over time of our fragrance-repellent we can now derive our objective stickiness equation and plug in our time-based viscosity equation.

Methods

In the objective stickiness equation section, we consider four main factors to construct our model: adhesion and cohesion are the main factors in determining stickiness, while viscosity and particle velocity determine how adhesion and cohesion translate into stickiness (Dunnewind et al., 2004; Gay, 2002). Thus, this section will be split into five parts: defining parameters & variables, deriving the equation by considering the contribution of adhesion & cohesion, deriving the equation by considering the contribution of viscosity, deriving the equation by considering the contribution of particle velocity, and assumptions on constants.

Defining Parameters & Variables

Stickiness Parameters: Cohesion, Adhesion, Viscosity, Particle Velocity

Stickiness(t) =f(A,C,V,Pv)

  • \(A\) = Adhesion
  • \(C\)= Cohesion
  • \(V\) = Viscosity
  • \(P\)v = Particle Velocity

Variables

  • \(S_0\)= initial stickiness due to cohesion (in kPa)
  • \(c\) = cohesion decay constant (in s-1)
  • \(k\)= adhesion decay constant (in s-1)
  • \(t\) = time (in seconds)
  • \(k_a\) = adhesion constant based on the solvent (no unit)
  • \(C_f\) = concentration of fragrance solute (in M)
  • \(v_p\)= velocity of the spray particles (in m/s)
  • \(τ\)= relaxation time (in seconds)

Deriving the Equation - Adhesion & Cohesion Contribution

We represent the combined contribution of adhesion & cohesion as

\[ A + C \]

Cohesion goes through decay over time (Gbaguidi et al., 2023), so we need to add a decay term c to the equation to represent the decay of the initial influence of adhesion and cohesion, which yields:

\[ A + C = (A + C)_0 e^{-c t} \]

Fragrance solvent concentration affects the strengths levels of both adhesion and cohesion (Diaz-Vecino et al., 2023). To accurately represent the influence of adhesion and cohesion it is crucial to consider the solvent concentration:

\[ A + C = (A + C)_0 e^{-c t} + k_a \cdot C_f e^{-k t} \]

\(k_a\) is a constant representing the increase in adhesion for each concentration unit of solvent, while \(C_f\) is the concentration of the fragrance. Adhesion goes through decay over time (Gbaguidi et al., 2023), so we need to add a decay term to the equation to represent the decay of the initial influence of adhesion and cohesion.

We then implement the stickiness variable \(S\) for ease of understanding, as well as to represent the idea that adhesion has zero contribution to initial stickiness (all the repellent molecules would be isolated with each other) without overrepresenting the importance of cohesion in determining stickiness.

\[ S = A + C \rightarrow S = S_0 e^{-c t} + k_a \cdot C_f e^{-k t} \]

Deriving the Equation - Particle Velocity Contribution

Particle velocity affects the stickiness of a solution (Sewalt et al., 2021;Diaz-Vecino et al., 2023). Particle velocity is the speed at which particles move at the moment of contact with a surface. \(τ\) is the relaxation time of a substance, which shows how quickly it responds to stress. We represent this relationship with:

\[ \left(1 + \frac{v_p}{\tau}\right) \]

At low particle velocities, the repellent molecules have more time to interact with the skin, resulting in greater adhesion due to ample time to facilitate deformation (the deformation of a substance leads to a greater contact area) and intermolecular bonding processes. At high particle velocities, the opposite is true, and particles can bounce off of the surface due to higher kinetic energy upon contact, resulting in lower adhesion (Sewalt et al., 2021;Diaz-Vecino et al., 2023). Knowing this, we can derive:

\[ S = S_0 e^{-c t} + k_a \cdot C_f e^{-k t} \cdot \left(1 + \frac{v_p}{\tau}\right) \]

Deriving the Equation - Viscosity Contribution

Viscosity affects both adhesion and cohesion, making it an important factor in determining stickiness (Papoutsi & Lianos, 1995). Viscosity is inversely proportional to adhesion, while being proportional to cohesion and generally proportional to the feeling of stickiness. We considered the previously derived time-based viscosity equation

\[ n(t) = n_0 \cdot \left(1 + k \cdot \left(\frac{V_c}{(V_{w0} + V_c - E \cdot t)}\right)^n\right) \]

which models the viscous behavior of our product over time. Implementing this into the Objective Stickiness Model, we get the final equation:

\[ \operatorname{Stickiness}(t) = \left(S_0 e^{-c t} + k_a C_f e^{-k t}\right) \cdot n_0 \cdot \left(1 + k \cdot \left(\frac{V_c}{(V_{w0} + V_c - E \cdot t)}\right)^n\right) \cdot \left(1 + \frac{v_p}{\tau}\right) \]

Assumptions

\(K_a\):

We assume Ka to be equal to 0.150. This assumption is based on the conventional values of the adhesion of water, which often falls within 0.100 J/m2 to 0.500 J/m2 (Papoutsi & Lianos, 1995). Based on the water-based composition of our repellent, and considering the strong hydrogen bonding exhibited between water molecules as well as the interference from the presence of minicells, a generally useful assumed value of the adhesion constant based on our solvent, water, is 0.150, which falls between the typical range of 0.100 J/m2 to 0.300 J/m2. Additionally, Ka’s purpose in the model is as a proportionality constant, therefore it does not have a unit.

\(c\):

We assume c to be equal to 0.200 s-1. This assumption is based on the conventional values of the cohesion of water after reviewing several studies, which often falls within 0.100 to 0.300 s-1 (Gbaguidi et al., 2023). Since our repellent is primarily water in composition, it is reasonable to use a cohesion decay value that represents water as a reference value to estimate the cohesion decay constant of our system. Due to the interference of minicells in our product, we assume the cohesion decay of water to be moderate, indicating a moderate cohesion decay constant value of 0.200 s-1.

\(k\):

We assume k to be equal to 0.100 s-1. This represents a moderate adhesion diminishing rate which makes our product’s adhesive properties similar to cosmetic fragrances (Gay, 2002). At an adhesion decay rate of 0.100 s-1, the repellent would stick for long enough to exert an effect while not adhering for an unrealistic amount of time.

τ:

We assume τ to be roughly 0.500 s. We make this assumption based on the behavior of cosmetic fragrances, which often exhibit similar relaxation time properties according to sensory evaluations.

Practical Analysis

To compare the stickiness of market repellents and BOROHMA, we used our Objective Stickiness Equation to predict and graph the total stickiness felt over a period of time for both products. For convenience, we estimated the cohesion and adhesion constants for the market repellent based on its viscosity difference from BOROHMA, as represented by differences in droplet sizes. To obtain precise and accurate measurements for droplet sizes, we used ImageJ, a precise image-measuring tool, to measure the diameters of droplets in pixels. We then obtained a ratio for the size difference of the market repellent and BOROHMA droplets and used this ratio to predict the new constants for the model when applied to market repellents.

Figure 7. Image of Market Repellent droplets in a spread test analyzed by ImageJ, where “Length” is measured in pixels and image scale is 1 centimeter = 280.029 pixels

Figure 8. Image of BOROHMA repellent droplets in a spread test analyzed by ImageJ, where “Length” is measured in pixels and the image scale is 1 centimeter = 200.090 pixels

Table 2. A table compiling the data obtained from analyzing the spread test results using ImageJ

From the data shown in Table 1, we obtained that the size-ratio between market repellent and BOROHMA droplet sizes is 1.320/0.874=1.510. Thereby, according to our proposed method for estimating constants for market repellents:


Table 3. A table displaying constants for Objective Stickiness Equation pertaining to BOROHMA and Market Repellent

We also assume \(S_{\text{0}}\) for BOROHMA is roughly 0.5 kPa, and the \(S_{\text{0}}\) for market repellents is roughly 15.0 kPa. Experimental procedures helped us determine Cf, yielding 0.4 OD minicell concentration for BOROHMA which is approximately 1.196M. We also assume the concentration in molarity ratio by effectiveness between BOROHMA and market repellents is 1:0.786 based on an AI-assisted estimation, meaning a market repellent concentration with an equivalent effect to 1.196M minicell concentration in BOROHMA is 0.942M. Implementing these values into the Objective Stickiness Equation for both products we get:

\[ S_b(t) = \left(0.5(0.82)^t + 0.1794(0.90)^t\right) \cdot \left(13.2 + \left(\frac{2.14 \times 10^4}{(150 - t)^{1.78}}\right)\right) \]

\[ S_m(t) = \left(15.0(0.88)^t + 0.0848(0.94)^t\right) \cdot \left(63.2 + \left(\frac{1.55 \times 10^5}{(150 - t)^{1.78}}\right)\right) \]

\(S_b(t)\) functions the stickiness behavior over time for BOROHMA and \(S_m(t)\) functions the stickiness behavior over time for market repellents. Using these functions we can obtain a Stickiness vs Time graph where the blue function represents the market repellent and the purple function represents BOROHMA:

Figure 9. A graph comparing the stickiness properties of market repellents and our insect-repellent fragrance by comparing integrals

From Figure 9, if we take the integral of the two functions, it is clear that over the same period of time (2 minutes) our product yields less of a sticky feeling than market repellents as shown with the equation

\[ \int_0^{120}(S_m(t) - S_b(t)) \, dt = 483.495 \]

The positive value supports the statement that our product offers a more comfortable user experience than market repellents as for the same effectiveness, less stickiness is felt over the same amount of time.

Results & Conclusion

Overall the results of this model prove that BOROMA solves a major problem that users face while using traditional oil-based repellents—stickiness and discomfort. Our water-based formula with the use of minicells is shown to not thicken and becomes less sticky over time. Our product remains comfortable even after five minutes with a viscosity of 2.026 cP as opposed to the estimated 4cP of the current oil-based repellents thus addressing the 75% of the users who complained about the stickiness of the current products. However, more importantly, BOROHMA has less initial stickiness compared to traditional repellents, which is 0.5 kPa against 15.0 kPa. This minimizes the chances of having to stick to the skin hence faster absorption and less irritation felt on the skin. Small droplet size and slower decay of cohesiveness make BOROHMA more effective than typical repellents in the short and long term within the framework of comfort. Ultimately, BOROHMA not only meets but exceeds user expectations for comfort and usability, making it a standout product in the mosquito repellent market. Its utilization of a water base positions it as the optimal choice for consumers seeking effective mosquito protection without compromising on comfort.

Limitations

While the Stickiness Metrics Framework provides a thorough and scientific approach to quantifying stickiness through viscosity, adhesion, and cohesion models, it is important to acknowledge that stickiness is a subjective experience. The level of stickiness thus may vary with variations in the skin thickness, the nature of conditions that the product is applied under, and individual tolerance for discomfort.

  1. Subjective Variability: While the model seeks to measure stickiness in terms of viscosity and behavior of particles, different users may feel a different level of discomfort because of their skin type or some factors such as humidity and sweat which change the feel of the stickiness of a repellent.
  2. Quantification Challenges: Stickiness has many parameters that depend on several factors that are not easy to measure to a great extent. While we have viscosity and droplet size which gives some indication of how something will feel, it doesn’t give any insight to the other sensations of the material despite being sticky and oily. It is challenging to quantify and represent them with reasonable reliability within a laboratory-style environment.
  3. Personal Bias: This must be taken into consideration though as the comfort is based on the level of skin temperature people may have some predisposed views about the texture or the type of repellents which may influence their comfort levels. Others may describe even a minimally sticky solution as irritating if the skin to which it is applied is normally dry.

  1. Diaz-Vecino, C., Rossi, E., Pollastri, S., Fries, A., Lemus, J., & Bonadonna, C. (2023). Insights into the sticking probability of volcanic ash particles from laboratory experiments. Scientific Reports, 13(1). https://doi.org/10.1038/s41598-023-47712-6
  2. Dunnewind, B., Janssen, A. M., van Vliet, T., & Weenen, H. (2004). Relative importance of cohesion and adhesion for sensory stickiness of semisolid foods. Journal of Texture Studies, 35(6), 603-620. https://doi.org/10.1111/j.1745-4603.2004.35512.x
  3. Gay, C. (2002). Stickiness—some fundamentals of adhesion. Integrative and Comparative Biology, 42(6), 1123–1126. https://doi.org/10.1093/icb/42.6.1123
  4. Gbaguidi, S. V., Odunlami, N. E. G., Zevounou, C., & Agbodjan, C. P. (2023). Variability of cohesion and angle of friction as a function of water content in fine soils in the Municipality of Lalo (Benin). World Journal of Advanced Research and Reviews, 20(1), 142–152. https://doi.org/10.30574/wjarr.2023.20.1.1993
  5. Papoutsi, D., & Lianos, P. (1995). Pyrene in mixed Titania-surfactant films made by hydrolysis of titanium isopropoxide in the presence of reversed micelles. Langmuir, 11(1), 1–4. https://doi.org/10.1021/la00001a001
  6. Sewalt, E. J., Zhang, F., Van Steijn, V., Van Ommen, J. R., & Meesters, G. M. (2020). Static and dynamic stickiness tests to measure particle stickiness. KONA Powder and Particle Journal, 38(0), 26–41. https://doi.org/10.14356/kona.2021017
  7. The influence of particles on suspension rheology | Anton Paar Wiki. (n.d.). Anton Paar. https://wiki.anton-paar.com/en/the-influence-of-particles-on-suspension-rheology/
  8. Xia, B., & Krueger, P. S. (2022). Rheology of particulate suspensions with non-Newtonian fluids in capillaries. Proceedings of the Royal Society A Mathematical Physical and Engineering Sciences, 478(2262). https://doi.org/10.1098/rspa.2021.0615
  9. von Fraunhofer, J. A. (2012). Adhesion and cohesion. International Journal of Dentistry, 2012, 951324. https://doi.org/10.1155/2012/951324




Market Adoption Model

Overview

Our product’s uniqueness comes from its dual-usage, combining insect-repellency with pleasant scents. However, there are very few similar products on the market that effectively perform both tasks well; realizing that our product has the potential to become the first successful solution of its kind, we wanted to create a predictive economic model to gain a better understanding of how our double-edged product would fare on the market.

Method

Introduction: Bass Diffusion Model

The Bass model is a well-established means of modeling the process of how new ideas and goods will be adopted and spread within a market or community; more precisely, it predicts the adoption curve of a new product by considering both innovators (the first people to adopt the new product without external influence) and imitators (people who adopt the product later who are influenced by earlier adopters) (Lenk and Rao, 1990). The number of new adopters as a function of time, denoted by \( N(t) \), is weighed by three parameters:

  • coefficient of innovation ( \( p \) ): the likelihood of the adoption of a product by innovators, the first group of consumers willing to try a new product or technology.
  • coefficient of imitation ( \( q \) ): the likelihood that someone will adopt the product due to the influence of the innovators.
  • market size ( \( m \) ): total number of potential adopters in the market.

Combining Parameters

Obtaining the coefficients of innovation and imitation with market data for an insect-repellent fragrance is extremely difficult due to the scarcity of similar products and legal sensitivities surrounding business data and information. It is, however, logical to assume the demographic of consumers that would be interested in an insect-repellent fragrance is the overlapping adopters of both the insect-repellent and fragrances industries. Still, we do not know which industry’s adopters will have a bigger impact on the overall adoption of our repellent-fragrance product, that’s why we devise the following weighted-average method to combine coefficients and market size:

Let \( p_r \) and \( q_r \) be the coefficients of innovation and imitation, respectively, for the repellent industry. Similarly, let \( p_f \) and \( q_f \) be the coefficients of innovation and imitation for the fragrance industry. A weighted average of \( p_r \) and \( p_f \), adjusted based on how much emphasis the product places on repellent versus fragrance, can be obtained. Likewise, a weighted average of \( q_r \) and \( q_f \), and \( m_r \) and \( m_f \), can be calculated:

\[ p_c = p_r \times w_r + p_f \times w_f \] \[ q_c = q_r \times w_r + q_f \times w_f \] \[ m_c = m_r \times w_r + m_f \times w_f \]

where \( p_c \) and \( q_c \) are the coefficients of innovation and imitation, respectively, of the repellent-fragrance industry. \( w_r \) and \( w_f \) are the weights representing the importance of each market (e.g., if the repellent function is more central to the product, \( w_r \) would be higher than \( w_f \)).

Obtaining Parameters

There are several methods to estimate the coefficients of the Bass model. Linear and nonlinear regression can be used if we have historical sales data for a product for a few periods. The simplest way to estimate is via nonlinear regression. Given at least four observations of \( N(t) \) we can use nonlinear regression to estimate parameter values (\( p_r \), \( q_r \), \( m_r \), \( p_f \), \( q_f \), \( m_f \)) (Srinivasan and Mason, 1986). After obtaining data obtained from Statista shown in Figure 1 & Figure 2 ("Fragrances - United States," 2024; "Mosquito repellent market value," 2023), we divided the market size in USD by the average price of each product to find the number of sales over time (Berr, 2016; "Perfume market analysis," 2020).


Figure 1. Market size of mosquito repellent products in the United States (in millions of USD)


Figure 2. Market size of the fragrances market in the United States (in billions of USD)

After regressing data for the repellent market, we found the coefficient of innovation to be \( 0.2 \cdot 10^{-10} \), the coefficient of imitation to be around \( 0.07673 \), and the market size to be around \( 2.319675 \times 10^9 \) (Figure 3). Similarly, we found the coefficient of innovation to be \( 0.3 \cdot 10^{-9} \), the coefficient of imitation to be around \( 0.01573 \), and the market size to be around \( 5.98762 \times 10^8 \) for the fragrance industry (Figure 4).


Figure 3. The bass diffusion plot for the repellent market


Figure 4. The bass diffusion plot for the fragrance market

Results & Conclusion

We created a simulation of the Bass model, with adjustable sliders for the weightings \( w_r \) and \( w_f \).

0.5
0.5

Based on our simulation, we found the optimal weighting to be when \( w_r \) was around 0.4 and \( w_f \) was around 0.6. Therefore, we decided that a heavier emphasis should be placed on the fragrance aspect of our project, resulting in the development of AroMatch. At the same time, the Bass model predicted a much higher number of total adopters for a combined repellent-fragrance market compared to each market considered separately. Thus, our entrepreneurship plans placed a heavy emphasis on the dual-functionality of BOROHMA. Based on the estimated number of total adopters, it can also be concluded that our product has great potential in a combined repellent-fragrance market and is worthy of being invested in and mass-produced in the future.

Limitations

Although our weighted average method is logical and should provide fairly reasonable results, it is not a typical way to utilize the coefficients of innovation and imitation. Moreover, there is a key limitation to the Bass diffusion model: it assumes consistent parameters over time, whereas, in reality, these parameters can alternate due to different factors such as investment and competition, meaning that the model cannot account for many external factors. The model also assumes that there is a homogeneous market, where every adopter has the same traits and behaviors. Realistically, markets have clients that range in terms of adoption behaviors (Bass et al., 1994). By dividing the total market size in USD by the average price of each product, we also failed to account for those who purchase a product more than once.

  1. Berr, J. (2016, May 27). For consumers, avoiding Zika is getting more expensive. CBS News. https://www.cbsnews.com/news/as-demand-for-bug-spray-raises-on-zika-fears-so-do-prices/
  2. Perfume market analysis 2019~2020 - OS Fragrance. OS Fragrance - Perfume Manufacturer, Supplier & Trading. (2020, September 22). https://www.osperfume.com/perfume-market-analysis.html
  3. Mosquito repellent market value U.S. 2025. Statista. (2023, March 24). https://www.statista.com/statistics/992924/us-mosquito-repellent-market-size/
  4. Fragrances - United States: Statista market forecast. Statista. (2024, September). https://www.statista.com/outlook/cmo/beauty-personal-care/fragrances/united-states
  5. Bass, Frank M.; Krishnan, Trichy V.; and Jain, Dipak C. (1994). Why the Bass model fits without decision variables. Marketing Science, 13(3), 204–223.
  6. Lenk, Peter J. and Rao, Ambar (1990). New Models from Old: Forecasting Product Adoption By Hierarchical Bayes Procedure. Marketing Science, 9(1), 42-53.
  7. Srinivasan, V., and Mason, Charlotte H. (1986). Nonlinear Least Squares Estimation of New Product Diffusion Models. Marketing Science, 5(2), 169-178.