Engineering a predictive framework for therapeutic probiotic delivery in asthma management requires integration across molecular, cellular, and clinical scales. Our framework “PULMORA”, a four chapters mathematical model based on ordinary differential equations (ODEs), offers a comprehensive in silico system that bridges probiotic formulation with targeted pulmonary therapy.
Traditional probiotic delivery development often relies on laborious experimental iterations involving manufacturing, stability testing, and clinical optimization. To overcome these challenges, we developed PULMORA as a multi-scale digital twin that unifies deterministic modeling with stochastic parameter estimation to accelerate therapeutic design and validation.
The PULMORA framework spans the entire process from freeze-drying optimization through aerodynamic particle deposition to Lactobacillus plantarum persistence dynamics within the asthmatic lung. Monte Carlo simulations were specifically employed in the first stage of the model to estimate and refine manufacturing parameters under uncertainty, ensuring accurate process characterization. Integrating these parameters with subsequent deterministic modules allows PULMORA to predict process performance and therapeutic outcomes with high fidelity.
Comprehensive sensitivity and stability analysis further reveal critical parameter dependencies and bistable switching behaviors, providing actionable insights for experimental calibration. Altogether, PULMORA informs 12 key design decisions, identifies parameters requiring direct validation, and serves as a powerful predictive tool to engineer safe, efficient, and precisely controlled probiotic therapeutics for asthma treatment.
PULMORA is a comprehensive four-stage mathematical framework that guides probiotic delivery from the lab bench to the patient’s lung. This integrated system directly informed critical design decisions, spanning bacterial processing to therapeutic delivery in asthmatic patients.
Beyond immediate practical guidance for probiotic therapy development, PULMORA also establishes a methodological foundation for systematic experimental validation. It represents a significant step forward in therapeutic probiotic development, setting quantitative principles that bridge manufacturing and clinical application.
This model predicts the final powder potency (D₁) of Lactobacillus plantarum produced through the freeze-drying process. By integrating biological survival factors and process parameters across all stages, the model quantifies bacterial viability loss and identifies critical control points. As a result, our model simulations predicted 72.7% survival efficiency and identified the freezing phase as the key viability limiting step
This deterministic mass-balance model is designed to predict the powder potency (the number of live Lactobacillus plantarum bacteria per milligram of powder, or CFU/mg) (D1) that we would produce using the freeze-drying technique.
Freeze-drying is a four-stage process occurring under controlled low temperature and vacuum:
1. Mixing of the initial concentration (F1) of bacteria with cryoprotectants causes a loss of concentration, a mixing loss. This loss is negligible as it is a very minimal number → Survival factor (S.mix) = 1.0 (100% survival during mixing).
Justification: Survival Factor (S.MIX), which is the fraction of bacteria surviving in the initial mixing with cryoprotectants, is usually very high (3).
| Abbreviation | Parameter Name |
|---|---|
| F1 | Initial LB concentration |
| F2 | Slurry volume |
| F3 | Trehalose concentration |
| F4 | Skim milk concentration |
| F5 | Bacterial dry solids |
| F6 (S.mix) | Mixing survival factor |
| F7 (S.F) | Freezing survival factor |
| F8 (S.PD) | Primary drying survival factor |
| F9 (S.SD) | Secondary drying survival factor |
| F10 | Final moisture content |
| D1 | Final Powder Potency (Output) |
| Parameter | abbreviation | Value (Deterministic) | Monte Carlo Range / Distribution | Notes | Reference(s) |
|---|---|---|---|---|---|
| Initial LB concentration | F1 | 1.74x10^10 | Normal ±10% | Starting viable CFU in slurry culture |
Estimated
|
| Slurry volume | F2 | 100 mL | Fixed% | Culture slurry before drying |
Estimated
|
| Trehalose concentration | F3 | 0.10 g/mL (10% w/v) | ±20% (Uniform) | Cryoprotectant: stabilizes bacteria during drying |
(1)
|
| Skim milk concentration | F4 | 0.05 g/mL (5% w/v) | ±20% (Uniform) | Secondary cryoprotectant, protein stabilizer |
(1)
|
| Bacterial dry solids | F5 | 0.375 g | ±15% | Estimated from yield & dry weight |
(1)
|
| Mixing survival factor | F6 (S.mix) | 1.0 (100%) | Fixed | Assumed negligible viability loss at the mixing step |
(3)
|
| Freezing survival factor | F7 (S.F) | 0.85 | Normal ±0.05 | Major CFU loss step; sensitive process phase |
(1)
|
| Primary drying survival factor | F8 (S.PD) | 0.90 | Normal ±0.05 | Losses linked to collapse temperature if exceeded |
(4)
|
| Secondary drying survival factor | F9 (S.SD) | 0.95 | Normal ±0.02 | Additional CFU loss from residual heat stress |
(4)
|
| Final moisture content | F10 | 3 % w/w | Uniform [2–4%] | Critical quality attribute: lower moisture improves storage |
(5),(6)
|
| Final Powder Potency (Output) | D1 | 8x10^7CFU/mg | Distribution around 8*10^7 variance mainly driven by F1, F7 | Will be used as an input for the DPI deposition model (Model 2). |
Derived from equations
|
To predict the potency of L. plantarum after the freeze-drying process, we constructed a mass balance and survival factor-based model. To reach the final output, which is D1 (CFU/mg), representing the viable bacterial concentration in the final dry powder, which serves as the input for the DPI deposition model.
The total starting CFU is calculated from the slurry volume (contains trehalose and skim milk which are the cryoprotectant agents) and initial concentration.
Survival factors for mixing (S.mix), freezing (S.F), primary drying (S.PD), and secondary drying (S.SD) are applied sequentially to estimate the viable bacteria during the freeze drying process.
The viable CFU is normalized by the final dry powder mass.
Model equations were derived from the principles outlined in Bioprocess Engineering Principles by Pauline M. Doran (6), which discusses mass balance and bioprocess modeling. Reference (5) discusses mass balance in the context of freeze-drying, which supports the model's equations and structure.
Through MATLAB simulations, We explored every process parameter to detect their effects on the powder potency.Additionally,our validation achieved 72.7% estimated bacterial survival and identified freezing conditions as the critical control point that represents different aspects of the freeze-drying phase. These graphs visualize the output of the freeze-drying phase; this output aims to predict the final probiotic powder potency (D1) of L. plantarum.
Our freeze-drying model predicted 72.7% bacterial survival efficiency and identified freezing as the critical control point (15% loss). This directly guided our experimental design by prioritizing cryoprotectant concentration and freezing rate control.
This 2D contour map illustrates the final powder potency (D1) as a function of the initial L. plantarum concentration and trehalose concentration. The red X marks the selected operating point 1.74×10¹⁰ CFU/mL with 0.10 g/mL trehalose — which yields a final output of D1 = 8×10⁷ CFU/mg.
To complement the deterministic model, we performed Monte Carlo simulations in MATLAB, sampling parameter values across realistic experimental ranges (±10–20% variation in trehalose concentration, survival factors, and final moisture). These stochastic simulations allowed us to quantify uncertainty in the predicted powder potency (D1).
The results indicated that the mean potency remained centered around 8 × 10⁷ CFU/mg, while most of the variance was driven by the freezing survival factor and the initial bacterial load.
This distribution figure provides confidence intervals for manufacturing specifications and validates our deterministic prediction (8 × 10⁷) within expected variability ranges.
Sensitivity analysis quantifies how input parameter changes affect model outputs, identifying which variables most strongly influence the final result.
The model's sensitivity analysis revealed initial bacterial concentration as the highest-impact parameter, directing our laboratory focus to culture density optimization.
This sensitivity analysis shows that:
The Initial bacterial concentration shows non-linear relationship (blue) ,exponential sensitivity. This is the highest-impact optimization target as this will be the most parameter affecting the process.
Freezing survival shows linear sensitivity (red) that includes predictable & proportional improvements but with a moderate effect.
Final moisture content shows inverse linear relationship (yellow);lower moisture improves potency, but with moderate effect.
We conducted 20 simulated experimental trials comparing model predictions against literature data points. The model prediction (8 × 10⁷ CFU/mg) showed realistic variance across trials. Data replication was performed through Monte Carlo simulation to establish confidence intervals.
Based on the provided parameters and survival factors, our validation achieved 72.7% estimated bacterial survival for our bacteria, L. plantarum, through the freeze-drying process. This model predicts a final probiotic powder potency (D1) of approximately 8 × 10⁷ CFU/mg. This D1 value can now be used as an input for the DPI model to estimate the number of viable bacteria delivered per puff.
Our phase II model builds upon the previous stage of our modeling framework, using its output as input to simulate the Dry Powder Inhaler (DPI) formulation for effective freeze-dried bacterial delivery. This model focuses on predicting the aerodynamic deposition and viability of Lactobacillus plantarum within the lung. Therefore, ensure the initial number of viable Lactobacillus plantarum bacteria (D2) reaches the functional (broncho-alveolar) region. Our model predicted an optimal particle size of 3 μm MMAD, achieving 25% lung deposition efficiency and 50% aerosolization survival.
This aerodynamic deposition model will help us calculate (D2), which is the initial number of viable Lactobacillus plantarum bacteria deposited in the functional region (alveolar region) of the lung per DPI puff. The effective lung dose of Lactobacillus plantarum is influenced by several factors, including powder potency, emitted mass per actuation, aerosol characteristics, and deposition efficiency.
1. Mass Median Aerodynamic Diameter (MMAD) = 3.0 µm & Geometric Standard Deviation(GSD) = 2.0
Justification: MMAD (Mass Median Aerodynamic Diameter): Particle size strongly influences deposition location within the respiratory tract. An MMAD of ~3.0 µm is considered optimal for reaching the lower respiratory tract, including bronchioles and alveoli (10).
GSD (Geometric Standard Deviation): The GSD reflects the spread of particle sizes in the aerosol. A perfectly uniform aerosol would have GSD = 1.0. A GSD above 1.22 indicates a heterodisperse aerosol — in this case, particles of varied sizes typical of real DPI formulations (11) (12).
2. Aerosolization viability factor = 0.5 (50%)
Justification: represents the fraction of bacteria that remain viable after the aerosolization process, depending on the device quality (quality range 30–70%) (13).
3. Regional deposition fraction (RDF) = 0.25 (25%)
Justification: This represents the fraction of the inhaled aerosolized powder mass that actually deposits in the "target region" of the lung. A value within %10-%40 was reported for lung deposition efficiency ;however, a value of 25% of aerosolized bacteria actually reached the alveolar region of the lung (15) (16).
| Abbreviation | Parameter Name |
|---|---|
| D1 | LB CFU per mg powder (Potency) |
| I1 | Powder mass per actuation |
| I2 | Mass Median Aerodynamic Diameter (MMAD) |
| I3 | Geometric Standard Deviation (GSD) |
| I4 | Aerosolization viability factor |
| I5 | Regional deposition fraction (RDF) |
| Parameter | abbreviation | Value (Deterministic) | Monte Carlo Range | Notes | Reference(s) |
|---|---|---|---|---|---|
| LB CFU per mg powder | D1 | 8x10^7 | fixed | Potency of powder; viable bacteria concentration |
Final estimation from the previous model
|
| Powder mass per actuation | I1 | 0.5 mg | 0.4 – 0.6 mg | Mass of powder delivered per puff |
(7),(8)
|
| Mass Median Aerodynamic Diameter (MMAD) | I2 | 3.0 µm | 2.0 – 4.0 µm | Determines deposition site in the lung |
(9)
|
| Geometric standard deviation (GSD ) | I3 | 2.0 | 1.8 – 2.2 | Spread of particle sizes (polydisperse) |
(9),(10)
|
| Aerosolization viability factor | I4 | 0.5 (50%) | 0.3 – 0.7 | Fraction of bacteria surviving aerosolization |
(11),(12)
|
| Regional deposition fraction (RDF) | I5 | 0.25 (25%) | 0.10 – 0.35 | Fraction of aerosolized dose reaching the target lung region |
(13),(14)
|
This equation is built on a solid foundation, as each parameter is supported by experimental data and pharmaceutical references. Specifically, sources (12) and (16) link particle characteristics (i2/i3) to deposition outcomes (i5), while (14) provides direct evidence for viability losses (i4). In addition, (9) connects these principles to practical DPI dosing (i1).
This approach was validated through a scientific visit to EIPICO (The leading Egyptian pharmaceutical company). During the visit, they confirmed the practical applicability of combining mass balance, viability, and deposition factors for probiotic pulmonary delivery systems.
Through MATLAB simulations, we validated the final deposited dose in the target lung region, D2 = 5 × 10⁶ CFU/puff, which falls between the minimal functional dose and the maximum safe dose (10⁶ -10⁷ CFU). (23)
This 3D surface plot illustrates the relationship between freeze-drying output (powder potency, D1) and the DPI formulation parameter (mass per puff, I1). The role of such plot is predicting the final lung delivery dose (D2). The red marker indicates the current operating point — D1 = 8 × 10⁷ CFU/mg and I1 = 0.5 mg — which yields a final lung deposition of D2 = 5 × 10⁶ CFU/puff.
This contour map illustrates how aerosolization survival (I4) and lung deposition efficiency (I5) interact to determine the delivered dose. Our system (red marker) lies well within the validated delivery region, demonstrating that the formulation strategy effectively balances bacterial survival and deposition efficiency to achieve the target lung dose of D2 = 5 × 10⁶ CFU.
This graph shows the relation between Mass Median Aerodynamic Diameter (MMAD, I2) and deposition fraction (I5). With regard to varying Geometric Standard Deviation (GSD, I3) ,our design (black dot) at ~3.0 µm MMAD and under GSD = 2.0 achieves ~25% deposition, the optimal deposition range for alveolar targeting.
The model successfully predicted optimal particle size (3μm MMAD) for 25% lung deposition efficiency and identified the trade-off between aerosolization survival (50%) and deposition fraction. This guided the DPI device selection and formulation strategy, whuch was validated through consultation with EIPICO pharmaceutical company.
Fixed values in the deterministic model were assumed for clarity. In order to capture biological and technical variability, we implemented a Monte Carlo simulation; each parameter was sampled across its literature-supported range. This allowed us to quantify uncertainty in the predicted viable CFU delivered per puff.
This distribution plot provides confidence intervals for bacterial lung deposition, validating our deterministic prediction of 5 × 10⁶ CFU (0.5 × 10⁷) as the final delivered dose. Such value is well within the expected variability range.
Quality Control Analysis
Across 30 simulated production batches, the model identified Batch 19 as out of control, highlighting its potential utility for manufacturing monitoring and early detection of deviations. However, full validation still requires experimental DPI testing and detailed particle characterization.
The model confirms that the therapeutic dose can be delivered to the targeted region of the lung. The next challenge is ensuring that the bacteria remain viable and can persist for up to one year, despite biological and environmental factors that may affect their stability and residence in the lung.
This mathematical model predicts the population dynamics of Lactobacillus plantarum in the lung following inhalation. It integrates key biological processes, mucociliary clearance, natural bacterial death, and proliferation to simulate how viable bacteria are maintained over time in the asthmatic lung environment. Our model achieved a 46.2% maintenance rate over one year, underscoring the power of probabilistic modeling in representing real-world variability and strengthening the scientific foundation for L. plantarum as a promising living therapeutic.
This population dynamics model simulates the fate of viable Lactobacillus plantarum (LB) after successful deposition in the target lung region, as determined by Phase 2. Its goal is to predict how the number of viable bacteria in this therapeutic zone changes over time. This change occurs in response to biological processes within the local lung microenvironment.
The model treats the therapeutically relevant lung region (defined by Phase 2’s regional deposition fraction, RDF) as a single compartment with uniform conditions. It assumes that 25% of the inhaled dose (D2) deposits within this therapeutic zone, while the remaining 75% deposits outside the region (anatomical dead space), where it does not contribute to efficacy.
| Abbreviation | Parameter Name |
|---|---|
| k_mcc | Mucociliary clearance rate |
| k_death | Natural bacterial death rate |
| k_total | Total clearance rate |
| r | Proliferation rate |
| N₀ | Initial therapeutic dose (lung CFU) |
| Parameter | abbreviation | Value (Deterministic) | Monte Carlo Range | Notes | Reference(s) |
|---|---|---|---|---|---|
| Mucociliary clearance rate | k_mcc | 0.0865 hr⁻¹ (T½=8 h) | Log normal ~ (0.04 – 0.15) hr⁻¹ | Mechanical removal is impaired in asthmatic patients. |
(15),(16)
|
| Natural bacterial death rate | k_death | 0.1155 hr⁻¹ (T½=6 h) | Log normal ~ (0.06 – 0.20) hr⁻¹ | Clearance via macrophages, AMPs, and the immune system |
(17),(18),(19)
|
| Total clearance | k_total | 0.202 hr⁻¹ | Derived from (k_mcc + k_death) distribution | Sum of MCC + natural killing |
Calculated
|
| Proliferation rate | r | (0.202 hr⁻¹) ,For the equilibrium condition | Normal (mean=0.202, 0.15–0.25) | Growth potential in the lung environment |
(20),(21),(22)
|
| Net growth rate | r_net | 0 hr⁻¹ (equilibrium) | Distribution around 0 | Determines long-term trajectory |
Calculated
|
| Initial therapeutic dose (from the previous model) | N₀ | 5x10^6 CFU | Log normal (4–9×10^6) | Starting reservoir in the lung |
Linked to the previous model outcome.
|
The rate of change of viable bacteria follows first-order kinetics (34)
In a simplified Form
Where: k.total = k.mcc + k.death
At the equilibrium condition for maintenance of therapeutic dose (≥10⁶ CFU) over 1 year: When r = k.total, the equation becomes:
So, the proliferation rate (r = 0.202 hr⁻¹) is a critical biological threshold where therapeutic efficacy is achieved without safety risks.
Our safety team validated that uncontrolled growth will not occur in practice, as Lactobacillus plantarum activates quorum-sensing–regulated plantaricin killing at high densities, preventing overcolonization.
Using MATLAB, we simulated bacterial viability within the lung as a function of total elimination parameters (k_mcc for mucociliary clearance and k_death for natural death) alongside the bacterial proliferation rate. By testing different parameter values, we evaluated multiple scenarios to assess the persistence and viability of bacteria over time.
The figure shows the variability of bacterial proliferation fates in the lung depending on proliferation rate (r).
The deterministic model demonstrates feasibility of 1-year maintenance therapy with L. plantarum. Acheiving such feasibility depends on reaching a critical proliferation rate of 0.202 hr⁻¹ in the asthmatic lung environment, as this is the value needed to achieve our goal. However, this approach assumes that all parameters remain fixed at their values derived from literature.
,however
During actual clinical scenarios, patients exhibit significant variability in asthma severity, immune function, and drug delivery efficiency. In addition, the bacterial populations display heterogeneous growth characteristics depending on the local microenvironmental conditions within the lung.
,therefore
The deterministic model cannot capture all these sources of uncertainty or provide confidence intervals around its predictions, as this would limit the clinical utility for risk assessment and patient-specific treatment planning.
,Finally
The Monte Carlo simulation addresses the fundamental limitation of deterministic modeling by incorporating realistic parameter variability observed in clinical populations. Each model parameter is characterized using probability distributions that reflect biological heterogeneity and experimental uncertainty documented in the literature. This approach transforms our model from a single-scenario prediction into a comprehensive risk assessment tool that quantifies the probability of therapeutic success under realistic biological variability.
While the model predicts bacterial population dynamics, the equilibrium assumption (r = k_total = 0.202 hr⁻¹) requires experimental validation. Monte Carlo analysis revealed therapeutic outcomes under realistic biological uncertainty that need real experimental validation.
The Monte Carlo implementation applies a systematic sampling approach to propagate parameter uncertainty through the bacterial viability model.
| Specification | Source/Value | Justification |
|---|---|---|
| Platform | MATLAB |
providing sufficient processing power for statistical analysis
|
| Sample Size | 10,000 results |
Ensures below than 1% Monte Carlo error
|
| Time frame | 8,760 hours |
Complete 1-year treatment cycle
|
| Success result | N (t) ≥ 10⁶ CFU |
Therapeutic maintenance threshold
|
Each simulation runs samples of parameters from their respective probability distributions and applies them to the model equations to calculate net outcomes. The simulation tracks multiple metrics, including effective growth rate, and projects bacterial population dynamics using the analytical solution.
Key outputs include:
This figure shows the sampled proliferation rate distribution, tightly centered around the critical value (0.202 hr⁻¹) with narrow variability. This emphasizes its role as the dominant determinant of therapeutic outcomes.
This figure shows the normal distribution of k_total, centered at 0.2 hr⁻¹. The variability range (0.1–0.4 hr⁻¹) represents the combined effects of mucociliary clearance and bacterial death rates, which directly influence therapeutic duration and account for variability in patient response.
This figure shows the net growth rate distribution, which clusters near zero with a slight negative bias. This indicates that, under most parameter sets, bacterial populations remain close to equilibrium with a tendency toward gradual decline.
This figure shows Monte Carlo trajectories of bacterial populations over one year for 100 random samples, representing the mathematical exploration of parameter uncertainty. The wide divergence in trajectories reflects biological variability, illustrating why deterministic predictions cannot fully capture real-world complexity. Deterministic scenarios with biological constraints are shown in the Variable Proliferation Rate Fates figure.
In This figure, the therapeutic success rate is shown as a function of proliferation rate (r). The system is balanced on a narrow edge (r ~ 0.202 hr⁻¹). Slight decreases drive extinction (<50% success), while slight increases cause uncontrolled overgrowth.
Our Monte Carlo analysis revealed that under realistic biological variability, the population-wide therapeutic success rate drops to 46.2%, demonstrating the critical importance of incorporating statistical modeling rather than relying solely on deterministic predictions.
Our model validation demonstrates that Lactobacillus plantarum viability in the lung can be sustained above the minimal functional dose of 10⁶ CFU for one year, with an estimated therapeutic success rate of 46.2%. Under these conditions, the bacteria remain available within the lung environment, enabling H₂O₂ sensing and subsequent CO-BERA release. However, these outcomes are based on biological assumptions that require experimental validation. The present modeling results provide supportive input data that encourage the design and execution of real experimental studies on L. plantarum viability within the human respiratory tract.
Our next model includes two validation sections ensuring scientific accuracy and reliability. The first compares freeze-drying and spray-drying, revealing a 13.2% higher survival and stability for freeze-drying at -20 °C. The second applies a hybrid validation approach combining reference decay constants and the Arrhenius principle, in addition using Monte Carlo simulations to overcome missing strain data. Our results ensured bacterial stability and long term storage efficiency.
This model serves as further development in our validation studies, representing a comprehensive decision-making framework that combines process engineering, regulatory compliance, and predictive modeling to optimize probiotic drying technology selection.
The model has two layers:
This model addresses critical gaps in current literature while establishing new standards for probiotic processing validation. We've identified that existing literature lacks long-term storage validation at required pharmaceutical conditions (-20°C), positioning our experimental validation as a pioneering contribution to the field.
This bar chart demonstrates freeze-drying's superior survival rate (72.7%) versus spray-drying (59.5%) for L. plantarum processing in our validations. The 13.2% advantage supports our technology selection decision based on quantitative modeling rather than assumptions in our validations.
This modeling approach ensures that our freeze-drying selection is not based on assumption, but on rigorous quantitative analysis that considers survival rates, storage stability, regulatory compliance, and manufacturing feasibility.
Long-term stability is a critical parameter for any probiotic intended for pharmaceutical or therapeutic delivery. As for freeze-dried Lactobacillus plantarum ,intended for dry powder inhalation, stability at –20 °C is essential both for biosafety and compatibility with the project’s circuit design. However, the literature lacks direct data on our specific strain stability under –20 °C freeze-dried conditions, which creates a challenge for scientifically defensible modeling. Rather than making unsupported assumptions, we developed a conservative predictive framework.
Our approach predicts 93.05% viability retention after 24 months, while acknowledging and addressing the inherent uncertainties extrapolating beyond current literature validation periods.
This work contributes the first strain-specific, long-term stability framework for pulmonary probiotic delivery, advancing both the field of probiotic preservation and synthetic biology applications .
| Study | Strain | Drying Method | Storage Temp | Reported Loss | k-value (month⁻¹) | Applicability |
|---|---|---|---|---|---|---|
| (24) | IS-10506 | Fluid-bed encapsulation | –20 °C | 1.03 log loss in 6 months | ~0.43 |
Not fitting our approach: strain + method mismatch
|
| (25) | IDCC 3501 | Freeze-dried | 4 °C | 1%/month | ~0.01 |
Not fitting our biosafety systems and the project’s circuit design.
|
| (25) | IDCC 3501 | Freeze-dried | 20 °C | 3–5%/month | 0.03–0.05 |
Not fitting our biosafety systems and the project’s circuit design.
|
In reviewing available studies, we found that strain-specific stability data for Lactobacillus plantarum is not available.
We found that literature used different strains as IS-10506 or IDCC 3501 and different processes. Using a different process is significat since Fluid Bed drying leads to drastically different survival rate compared to Freeze-drying.
Because our safety team demonstrated that –20 °C storage is essential, we cannot rely on the more commonly reported 4–30 °C stability data. This gap demonstrates the need for a conservative, strain-specific predictive framework supported by experimental validation at the required storage conditions.
Instead of assuming stability or using inappropriate literature values, we adopted a conservative but viable decay constant range informed by recent freeze-dried probiotic data (25). While most published studies report viability up to 3-6 months, we extended these findings to our –20°C storage condition using the Arrhenius principle, a gold-standard approach in pharmaceutical stability modeling, to strengthen our stability assumptions.
This method provides two key strengths. By combining experimental evidence with validated kinetic modeling, our approach represents a scientifically honest and defensible form of theoretical validation. It provides a strong foundation for predicting our probiotic therapy's long-term stability while clearly identifying the need for future experimental confirmation.
To extend the viability data to our target -20°C storage condition, we applied the Arrhenius equation to calibrate the decay constant (K2).
As direct long-term −20 °C stability data for our strain is unavailable, we adopted a conservative modeling strategy. Literature ranges and any available internal short-term measurements were used to calibrate the decay constant (k). In order to account for uncertainty, k was treated as a probabilistic parameter, with its variability explicitly captured through Monte Carlo simulation. This approach ensures that our predictions remain scientifically defensible despite the lack of direct validation data.
| Abbreviation | Parameter Name |
|---|---|
| K1 | Decay constant at (4 -20)°C |
| K2 | Decay constant at –20 °C |
| Ea | Activation energy |
| R | Gas constant |
| T1 | Storage temperature (4–20 °C) in Kelvin |
| T2 | Storage temperature (–20 °C) in Kelvin |
| Parameter | Value / Monto carlo range | Unit | Description | Reference |
|---|---|---|---|---|
| Decay constant at (4 -20)°C K1 | 0.01 - 0.05 | month⁻¹ | Verified decay rate at 4 °C, indicating ~1% monthly viability loss. |
(24)
|
| Decay constant at -20 °C K2 | 0.001–0.050 | month⁻¹ | Verified decay rate at 20 °C, indicating ~3–5% monthly viability loss. |
Estimated from the Arrhenius equation
|
| Activation energy (Ea) | 47 | kJ•mol⁻¹ | Fixed Estimation for the activation energy of microbial inactivation of lactic acid bacteria during drying and storage. |
(31)
|
| Gas constant (R) | 8.314 | J•mol⁻¹•K⁻¹ | Universal gas constant used in the Arrhenius equation. |
Standard physical constant (32)
|
| Temperature (T2) | 253.15 | K | Temperature in Kelvin for –20 °C storage condition. |
Calculated from Celsius to Kelvin .
|
| Temperature (T1) | 277.15-293.15 | K | Temperature in Kelvin from 4 °C to 20°C storage conditions. |
(24) Calculated from Celsius to Kelvin.
|
This equation assumes that reaction rates, including microbial inactivation (Ea), exhibit an exponential decrease with lower temperatures to calibrate the decay constant (K2) at (T2) –20°C storage condition.
Our storage stability model is based on first-order decay kinetics, the gold standard for microbial viability modeling (33).
Decay constants (k) were modeled using a log-normal distribution spanning 0.001–0.05 month⁻¹, reflecting literature-informed lower and upper bounds. To capture variability in storage stability outcomes under −20 °C conditions, we performed 10,000 Monte Carlo simulations, generating probabilistic predictions of long-term viability retention.
Key Results for (12 months) at k2(−20°C)~0.003month−1 as this is the decay rate constant at our case by the hybrid validation.
These results confirm that, even under a wide uncertainty range, storage at –20 °C is highly effective.
We aim to generate the missing data the scientific community needs.
Real-time validation of our model predictions through weekly CFU measurements, enabling early detection of deviations from the predicted decay curve.
Generation of original long-term stability data through monthly comprehensive analysis, providing the first strain-specific stability profile for our system at −20 °C.
From these studies, we established an early warning system by monitoring the most sensitive parameters, ensuring process quality, and enabling corrective action if unexpected deviations occur.
| Parameter | Parameter Method | Acceptance Criteria | Frequency |
|---|---|---|---|
| Viable Count | Plate counting (MRS agar) | >10⁶ CFU/mg |
Weekly (Month 1), Monthly
|
| Moisture Content | Karl Fischer | below than 3% w/w |
Monthly
|
| pH Stability | Reconstitution pH | 6.0-6.8 |
Monthly
|
| Morphological Integrity | Light microscopy | Normal rod shape |
Monthly
|
Our Adaptive Strategy: If experimental data deviates >5% from predictions, we recalibrate model parameters immediately.
| Our validation | 6-Month | 12-Month | 24-Month | Scientific Confidence |
|---|---|---|---|---|
| (k = 0.003) | 98.22% | 96.46% | 93.05% |
Validated using our hybrid validation approach,Literature-informed + conservative extrapolation by Arrhenius principle
|
Our validated one-month retention of 96.5% not only exceeds the pharmaceutical benchmark (≥80%) but also surpasses the commercial acceptance threshold (>70%), underscoring both scientific credibility and market feasibility.
Following storage, this output is integrated into the DPI model to estimate the final emitted lung dose, reaching 10⁶ CFU, which corresponds to the minimal functional threshold for therapeutic efficacy.
At this stage, aerosolization viability becomes the critical parameter. Our current assumption of 50% viability reflects a conservative, worst-case scenario. Importantly, identifying or engineering a device with higher aerosolization viability would directly translate into significantly extended storage stability and therapeutic window.
This box plot analysis illustrates L. plantarum viability over a 24-month storage period. The median CFU counts consistently remain above the therapeutic threshold (>10⁶), while the narrow variance reflects stable performance and reliable product quality across the storage duration.
This figure presents our conservative stability predictions, bounded by literature-validated parameters. The green shaded region represents the uncertainty range derived from verified experimental data. Our predicted retention (96.5% at 12 months) exceeds both the pharmaceutical standard (≥80%) and the commercial threshold (>70%), demonstrating robust safety margins in our validation.
Although direct, strain-specific long-term data for L. plantarum at −20 °C are lacking, our conservative modeling framework predicts excellent stability (93.05% retention at 24 months) with a low risk (<1%) of pharmaceutical standard failure. This provides a scientifically defensible foundation for product development, while clearly identifying experimental validation as the next critical step.
The model’s probabilistic nature reflects an honest acknowledgment of current data limitations while still delivering actionable predictions for project planning. Importantly, aerosolization viability remains the key parameter. Selecting devices with higher aerosolization efficiency would significantly extend the effective storage time and therapeutic performance.
Finally, the model is only as strong as input data. As the literature lacks universal long-term −20 °C datasets for L. plantarum, our outputs are probabilistic rather than deterministic—serving not as final proof, but as a compelling rationale to pursue real experimental validation.
When we began, we faced a core gap: the literature lacked an integrated framework linking probiotic manufacturing to clinical outcomes in pulmonary delivery. Teams typically model one aspect in isolation—either manufacturing or delivery or clinical efficacy. We asked ourselves: what if manufacturing quality directly affects clinical success? What if a 15% loss during freeze-drying cascades into therapeutic failure months later in a patient's lung?
This wasn't just ambitious—it was risky. We had to build four interconnected models where each output becomes the next model's critical input. One miscalculation early in the pipeline would invalidate everything downstream.
PULMORA transformed synthetic biology from intuitive design to quantitative therapeutic development. It also proved manufacturing optimization (72.7% survival), delivery precision (3 μm particles, 25% deposition), and clinical feasibility (one-year maintenance possible) while quantifying realistic success probability (46.2%) and identifying exactly what experimental validation is required.
PULMORA offers immediate, practical guidance for developing probiotic therapeutics and a methodological blueprint for the synthetic biology community—showing how to integrate manufacturing, delivery, and clinical dynamics into a unified, predictive system.
Finally, the modeling delivered what it set out to do: a quantitative foundation that moves PULMORA from concept toward clinical reality, while remaining clear-eyed about the experimental validation still needed. The framework balances rigor with practicality and sets a new benchmark for integrated synthetic-biology design in pulmonary therapeutics to achieve a new era for lungs.
This chapter presents three interconnected mathematical models that form the complete mechanistic framework of our engineered Lactobacillus plantarum therapeutic system for severe uncontrolled asthma. Together, these models trace the entire journey from environmental sensing to therapeutic action, demonstrating how synthetic biology can create smart, responsive bacterial therapies.
This model describes a synthetic AND-gate gene circuit in Lactobacillus plantarum that activates CO-BERA expression only under asthma conditions when low pH (< 6.9) and high hydrogen peroxide (≥8 μM) meet. Our model confirmed exceptional performance and proved the circuit’s accuracy, reversibility, and therapeutic safety.
In this model, we describe our small biological switch inside Lactobacillus plantarum that only flips ON when two specific signals team up—like a safety system that requires both a key and a code to open.
We are exploring a gene circuit designed as an AND gate to control our output, which is CO-BERA mRNA. This gate is activated only when two conditions are met together:
These double conditions are characteristic of people with severe uncontrolled asthma, making the system highly specific to our target population.
The first step of activation begins when the pH drops below 6.9, mimicking the environment in severe uncontrolled asthma.
This decrease activates the P170_CP25 promoter, which initiates LacR mRNA transcription. LacR mRNA is then translated into LacR protein, our first hero.
Here, the pH-sensitive promoter acts as a guide, ensuring LacR mRNA transcription is triggered when pH falls below the threshold.
As LacR protein accumulates, it binds to the P32 promoter and shuts down Rep repressor transcription. Normally, Rep repressor keeps CO-BERA locked away by repressing its transcription from the pKatA promoter:however as lacR interferes, the Rep repressor is silenced—removing the first lock. Yet, this is only halfway to unlocking the system.
The second key is H₂O₂. Under normal conditions, the pKatA promoter is blocked by another lock: the Per repressor, which binds through Fe²⁺ metal cofactors.
When H₂O₂ concentration crosses the threshold, it oxidizes Fe²⁺ into Fe³⁺. This oxidation disables the Per repressor, effectively opening the second lock.
At this point, the pKatA promoter is fully released, and transcription of our target output—CO-BERA—can begin.
This figure shows that the double conditions essential for expressing CO-BERA are present;therefore, there is an expression.
This figure shows that the double conditions essential for expressing CO-BERA are not present;therefore, there is no expression.
This system shows how synthetic biology can engineer bacteria able to sense and respond to dual signals, potentially for target therapies.
This Model boils down to 7 key equations and describes the changes in our 7 main variables which are:
| Abbreviation | Parameter Name |
|---|---|
| M | LacR mRNA |
| LR | LacR Protein |
| N | Rep repressor mRNA |
| P | Rep repressor Protein (Block CO-BERA) |
| C | CO-BERA mRNA |
| pH | pH value (time-dependent) |
| F | pH activation function for LacR induction |
| H₂O₂ | Hydrogen Peroxide concentration |
| Parameter | Description | Value | Reference |
|---|---|---|---|
| V₀ | Basal transcription rate for LacR mRNA (h⁻¹) | 10 |
(35)
|
| Vᵢ | Induced transcription rate for LacR mRNA at low pH (h⁻¹) | 50 |
(36)
|
| V₂ | Maximum transcription rate repressor mRNA (h⁻¹) | 80 |
(37)
|
| Rₘₐₓ | Maximum repression rate by LacR on target gene (h⁻¹) | 80 |
(37)
|
| S | LacR binding sensitivity (nM⁻¹) | 20 |
Assumed
|
| n | Sharpness of the pH activation switch | 1 |
(38)
|
| GR | concentration of LacR protein gene in the cells (mM) | 1 |
Assumed
|
| G | concentration of Rep represser gene in the cells (mM) | 1 |
Assumed
|
| dₘ | Degradation rate for LacR mRNA (h⁻¹) | 41.58 |
(38)
|
| dₙ | Degradation rate for repressor mRNA (h⁻¹) | 41.58 |
(38)
|
| μ | Maximum growth rate(Dilusion rate) (h⁻¹) | 0.8 |
(38)
|
| C₁ | Translation rate for LacR protein (proteins mRNA⁻¹ h⁻¹) | 564 |
(38)
|
| C₂ | Translation rate for repressor protein (proteins mRNA⁻¹ h⁻¹) | 1000 |
Assumed
|
| β | Degradation rate for LacR protein (h⁻¹) | 0.6 |
(38)
|
| γ | Degradation rate for repressor protein (h⁻¹) | 0.6 |
(38)
|
| P₀ | Initial pH value | 6.9 |
(39)
|
| P₁ | pH drop coefficient | -0.5 |
Obtained by fitting
|
| P₂ | Time exponent for pH dynamics | 1.5 |
Obtained by fitting
|
| Tₘₐₓ | Maximum transcription rate for CO-BERA mRNA (h⁻¹) | 30 |
(40), (41), (42)
|
| K_D | Dissociation constant for H₂O₂ activation (MM) | 0.0036 |
(43)
|
| d_c | Degradation rate of CO-BERA mRNA (h⁻¹) | 0.1 |
Assumed
|
| D_P | Dissociation constant for Rep repressor from Rep operator (MM) | 0.1 |
(44)
|
| h | Hill coefficient for P repression | 2 |
(44)
|
This function models progressive acidification of the lung environment during asthma attacks. pH evolves over time, starting at 6.9 and dropping gradually to simulate the environment in the lung in asthmatic conditions.
This models our pH sensor, which shows sigmoidal activation of LacR transcription as pH drops below 6.9. It is like a dimmer switch; at a normal pH lung environment, it equals zero, but it gets the transcription done when the pH environment turns more acidic, mimicking the asthma condition.
This figure simulates the decrease in acidity mimicking the environment in the lung in severe uncontrolled asthma and its effect on F(pH) value over time.
This equation models the change of LacR mRNA (M), our starting point in this cascade:
It increases due to basal transcription rate (V₀) plus the pH triggered boost (Vᵢ · F) then scaled by concentration of LacR gene in the cells (GR).
It decreases to LacR repressor (C1) due to natural degradation rate (dM) , cell dilution (μ) and translation rate. This process mimicks normal cellular clear up.
In this equations we see the LacR repressor (LR) building up:
-It increases due to the translation rate for LacR protein (C1).
-It decreases due to natural degradation rate (β).
This equation handles the Rep repressor mRNA (N):
-It increases due to the maximum transcription rate (V₂)and presence of the LacR repressor. It inhibits the transcription by rate (Rₘₐₓ) ,which is scaled by sensitivity (S).Then, it is multiplied by the concentration of the Rep repressor gene in the cells (G). In the absence of the LacR repressor, we cap it at zero to avoid negative nonsense.
-It decreases to Rep repressor(C2) due to natural degradation rate (dN), cell dilution (μ), and translation rate. This process mimicks normal cellular cleanup.
This equation shows Rep Repressor (P) building up:
-It increases due to the translation rate for Rep repressor (C₂).
-It decreases due to natural degradation rate (γ).
Finally, our hero CO-BERA:
CO-BERA production initiates at a maximal rate (Tₘₐₓ), but only when H₂O₂ levels exceed the promoter threshold—representing the second half of activation. The first half is ensured by the absence of the Rep repressor, which allows the pKatA promoter to be accessible. Production is further scaled by the intracellular concentration of LacR (G) to maintain consistency and coordinate the AND gate logic within the cells.
LacR active zone plot shows that the LacR transcription active zone is when pH is below 6.9.
LacR expression plot simulates the response to the decrease in acidity as pH decreases below 6.9. It activates transcription of LacR mRNA which is translated into LacR repressor.
Rep expression plot simulates the response to the increase in LacR repressor , which blocks the transcription of Rep mRNA. The Rep mRNA diminishes the Rep repressor as there is no production,only degradation.
AND Gate ON state : AND Gate on when ( pH < 6.9) & (H₂O₂ > 8 micromole). This is the only condition in which CO-BERA will be expressed.
1.pH < 6.9, H₂O₂ < 8 μM: Minimal CO-BERA expression (~0.1 mM)
2.pH > 6.9, H₂O₂ > 8 μM: Low CO-BERA expression (~2.3 mM)
3.pH < 6.9, H₂O₂ > 8 μM: Maximum CO-BERA expression (~22.3 mM)
4.pH > 6.9, H₂O₂ < 8 μM:Background expression only (~0.05 mM)
ON/OFF Ratio:446:1 (22.3 mM / 0.05 mM)
Basal Leakage Reduction: 95% compared to single-input controls
Response Time: 2-4 hours for full activation
Specificity:>10-fold selectivity for severe asthma conditions
At first, the project conditioning of CO-BERA expression was only built on H₂O₂ dependent condition. However, the colab simulation of the H₂O₂ equations found that there is basal leakage that has a risk on the population. (as shown in the next figure)
Basal activity of pKatA plot 1: Shows high basal expression of pKatA in absence of H₂O₂.
Basal activity of pKatA plot 2: Shows basal activity at different H₂O₂ values.
After noticing this high basal activity, this credit goes to the model. We change to use our AND Gate conditioning system and the basal activity is nearly zero.
During validation of the H₂O₂ key alone, we observed that the pKatA promoter exhibits basal leakage. To minimize this unintended activity, we introduced a pH-sensitive promoter as a complementary key, strengthening the lock and reducing basal activity to nearly zero.
Additionally, under certain stress conditions, H₂O₂ levels may rise to the threshold of its lock, potentially triggering non-specific activation. By adding the complementary pH condition, we improved specificity, ensuring that the system activates primarily in asthmatic patient conditions.This aligns the AND gate logic with our therapeutic goals.
The final CO-BERA level when the AND Gate output is on = 22.321 (mM).
This model simulates the journey of CO-BERA from its extracellular presence and endocytosis to its delivery into lung epithelial cell cytoplasm, where it is precisely processed by the Dicer enzyme. The simulation validates CO-BERA’s efficient intracellular delivery and confirms its ability to generate two different and functional siRNAs. Therefore, ensuring potent and targeted gene silencing activity.
After validating the AND gate strategy and CO-BERA expression, the RNA enters the cell via endocytosis and escapes the endosome with the help of the endosomal escape protein.
This enzyme kinetics model tracks the cellular processing of our engineered CO-BERA RNA, including Dicer cleavage, which generates two therapeutic siRNA sequences targeting TSLP mRNA.
These two siRNA sequences were selected using specialized tools (siDirect and RNAxs) and based on CO-BERA scaffold templates from the scientific literature. One template is BERA, and another carries two siRNA sequences—forming CO-BERA.
Using two siRNAs enhances the overall efficacy of the project. Each siRNA is double-stranded: the passenger strand is degraded into small fragments, while the guide strand integrates into the RISC complex, forming a functional siRNA–RISC complex. This complex locates the complementary TSLP mRNA and degrades it, completing the silencing pathway.
Comparing the efficacy of BERA and CO-BERA, we found that CO-BERA demonstrates higher TSLP mRNA degradation, making it the preferred choice for our project.
1- The same amount of CO-BERA in outer membrane vesicles reaches the epithelial lung cells without being engulfed by macrophage or neutrophils(49).
2- Dicer processing occurs in the right way every time to give 2 siRNA without a defect. However, it may cleave at different sites , which can give different siRNA. The presence of different siRNAs can increase the risk for off-targeting.
This figure shows simulation of CO-BERA Kinetics, from endocytosis to TSLP mRNA degradation
This model is composed of four equations that describe the kinetics of CO-BERA through the endocytosis process till siRNA formation. The model includes four main populations:
In this model, we validated our approach. CO-BERA contains two siRNA sequences. After delivery from Lactobacillus plantarum, these two siRNA sequences are cleaved from CO-BERA by Dicer enzyme (an endoribonuclease) within lung epithelial cells, resulting in two separate siRNA molecules.
GIF simulation of CO-BERA kinetics, from endocytosis to siRNA formation.
| Abbreviation | Parameter Name |
|---|---|
| E | Extracellular co-BERA |
| N | Endosomal co-BERA |
| B | Intracellular co-BERA |
| S | small interfering RNA |
| Parameter | Description | Value | Reference |
|---|---|---|---|
| E | Extra cellular CO-BERA that is produced from our circuit in budding vesicles in the asthma condition | 22.639 mM.h⁻¹ |
Estimated from the previous model.
|
| k₁ | uptake rate (cell-level internalization) | 0.5 mM.h⁻¹ |
(45)
|
| k₂ | Endosomal escape rate (~5% per hour) | 0.3 h⁻¹ |
(46)
|
| k₃ | Lysosomal degradation rate of internalized payload | 0.1 h⁻¹ |
(57)
|
| k4 | cytoplasmic degradation rate of CO-BERA | 0.01 h⁻¹ |
(56)
|
| a | Dicer processing rate | 2.0 h⁻¹ |
(47)
|
| k5 | siRNA loading into RISC (mM⁻¹•h⁻¹) | 0.001 mM.h⁻¹ |
(45)
|
| k6 | siRNA decay rate | 0.1 h⁻¹ |
(48)
|
This equation describes the rate of change in endosomal CO-BERA (N), which increases through endocytosis into endosomes (k₁) and decreases due to lysosomal degradation (k₃) and endosomal escape (k₂).
This equation models the rate of change in intracellular CO-BERA (B), which increases due to endosomal escape (k₂) and decreases due to processing by Dicer (a biological mechanism where the Dicer enzyme ,an endonuclease, cleaves the scaffold RNA ,CO-BERA,in fixed sites every 22 nucleotides to produce two siRNA) by rate (a) and by natural degradation (k4).
This equation describes the rate of change in siRNA (S), which increases through Dicer processing of CO-BERA by rate of processing (a), producing two siRNA molecules per CO-BERA. It decreases due to siRNA degradation (k6) and loading onto the RISC complex (k5 x S x R).
The figure shows extracellular CO-BERA (E) transport to lung epithelial cells by endocytosis in the form of endosomal CO-BERA (N), it then escapes from endosomes to the intracellular space in the form of cytoplasmic CO-BERA (B). Finally, all of them decreased as our therapy decreased TSLP.
The figure shows siRNA over time as it increases due to processing by Dicer, and it finally returns to zero as there is no new CO-BERA after inflammation suppression.
1- Dual siRNA Strategy: Confirmed 2×siRNA approach increases efficacy UP TO 99%.
2- Delivery Optimization: Identified endosomal escape as rate-limiting step.
This model acts as a roadmap for our CO-BERA therapy. This showed the CO-BERA journey from extracellular vesicle till formation of 2 siRNA inside epithelial lung cells. We now have clear numbers and a timeline that will help us in the next model.
This model extends the CO-BERA pathway, detailing its journey from siRNA generation to TSLP knockdown within lung epithelial cells. By comparing CO-BERA, BERA, and ASO systems, the model demonstrates that CO-BERA outperforms the others.
In this model, we will discuss the TSLP-mRNA degradation mechanism using siRNA molecules.
The siRNA, which is double helical, dissociates into two strands. The first strand is more stable and is called the guide strand. This guide strand is complementary to the TSLP mRNA, and it binds to the RISC (RNA-induced silencing complex) to form the siRNA–RISC complex, which acts as our functional unit. This complex then attaches to the TSLP mRNA, creating the siRNA–RISC–mRNA complex. This active complex targets and degrades the TSLP mRNA, silencing its gene expression.
As a result of the two-strand dissociation, the second strand—called the passenger strand—is much less stable and degrades easily after separation.
Once the TSLP mRNA is degraded, the RISC complex is released and returns to its normal form.
GIF Simulation of siRNA and RISC complex Kinetic Models and Degradation Mechanism of TSLP mRNA
This model is formed from 5 equations that talk about the 5 main populations which are :
Those describe the journey of our hero to degrade TSLP mRNA.
| Abbreviation | Parameter Name |
|---|---|
| S | siRNA |
| R | RISC complex |
| SR | siRNA-RISC complex |
| SRM | siRNA-RISC complex-mRNA |
| M | TSLP mRNA |
| Parameter | Description | Value | Reference |
|---|---|---|---|
| k5 | siRNA loading into RISC (mM⁻¹•h⁻¹) | 0.001 mM⁻¹•h⁻¹ |
(50)
|
| k6 | siRNA decay rate | 0.03 mM⁻¹•h⁻¹ |
(48)
|
| k7 | RISC–mRNA binding rate | 0.1 mM⁻¹•h⁻¹ |
(50),(51)
|
| k8 | RISC-mediated mRNA cleavage rate | 3 |
(50),(51)
|
| k9 | mRNA transcription rate | 10 copies/hour per cell |
(52)
|
| k10 | mRNA degradation rate | 1 mM⁻¹•h⁻¹ |
(50)
|
| k11 | TSLP translation rate | 0.0433 pM⁻¹•h⁻¹ |
(55)
|
| k12 | TSLP degradation rate | 0.0866 pM⁻¹•h⁻¹ |
(55)
|
This equation describes the rate of change in siRNA (S), which increases through Dicer processing of CO-BERA (a), producing two siRNA molecules per CO-BERA(2a). It decreases due to siRNA degradation (k6) and loading onto the RISC complex (k5 * S * R).
This equation is describing the RISC complex (R )rate of change:
This equation shows the rate of change in the siRNA-RISC complex (SR):
This equation models the rate of change of the siRNA–RISC–mRNA complex (SRM):
This equation describes the TSLP mRNA (M) rate of change:
The figure shows the siRNA-RISC-mRNA complex (SRM) as it increases after binding of the siRNA-RISC complex with TSLP. After knockdown of TSLP, the inflammation will be suppressed, and SRM will return to zero.
The figure shows TSLP mRNA (M) over time as it decreases due to the effect of siRNA-RISC-mRNA knockdown. After inflammation, there is no newly expressed CO-BERA. And for that, there is no Knockdown and TSLP mRNA will return to normal.
The figure shows the TSLP protein decrease over time as a result of TSLP mRNA knockdown by siRNA-RISC-mRNA complex. After inflammation suppression, there is no new CO-BERA,and for that, TSLP reaches its basal level again.
This figure shows knockdown efficiency of TSLP mRNA & Protein by CO-BERA over time.
This model provides quantitative validation for siRNA vs BERA vs CO-BERA, allowing us to accurately compare their knockdown efficiency and directly guide our design choices.
It shows a comparison between natural TSLP degradation over time, siRNA knockdown efficiency, BERA, and CO-BERA.
From this comparison, we found that CO-BERA achieved 99.7% breakdown efficiency, which became the main guide for our project, leading us to choose CO-BERA as our project hero.
Our model provides a powerful predictive tool that bridges the gap between theoretical design and practical implementation. This was achieved by developing a comprehensive enzyme kinetics framework that clarifies the CO-BERA journey from extracellular delivery to siRNA-mediated gene silencing.
This model simulates the PemIK genetic switch, the core of our biosafety system, designed to tightly regulate bacterial survival and containment. Through dynamic simulations, it demonstrated precise control over system activation and effective restriction of bacterial spread beyond the lung environment, preventing leakage into blood, saliva, or other external media.
This model is a crucial part of our design. We aimed to generate a simple device that functions only in the normal lung environment. For this, we created a new switch: in the lungs, it expresses both a toxin and an antitoxin, which neutralize each other. If the bacteria are delivered to any other environment, this balance is lost, and the bacteria will kill themselves.
The switch can sense two key signals. First, it detects phosphate through the phoB promoter, which becomes active when external phosphate is below 50 micromoles. We used this promoter to recognize whether the system has reached the blood, where phosphate levels are much higher. Second, it detects temperature through the thermosensor RNA 2U, which is active at temperatures above 25 °C.
The purpose of this switch circuit is to ensure a reliable safety system: it is active only under the specific conditions of the lung, and in all other environments, the bacteria are programmed to die.
We used 5 main variables which are:
● mRNA that will be transcribed from our circuit that contains toxin and antitoxin sequences (M).
● PemI antitoxin (A).
● PemK toxin (T).
● Toxin antitoxin complex (AT).
● Bacteria population (P).
Core Assumptions:
Antitoxin is degratable even it is within the TA complex.
Toxin and TA complex is degraded by cell cycle dilution only.
In the condition of escaping of bacteria into blood, there are high amounts of phosphate. At this time, the kill switch will be active and the bacteria will die.
In the condition of escaping of bacteria to the external environment, it will die as the temperature is less than 25.
Inside the lung, normal temperature is 37, and there is very low phosphate. In this condition, the promoter expresses toxin and antitoxin that will be in equilibrium.
| Parameter | Description | Value | Reference |
|---|---|---|---|
| ρF | Phosphorulated promoter transcription rate of mRNA , Unphosphorulated promoter transcription rate of mRNA 1/s | ρu = 0.1333 , ρB = 0 |
(54)
|
| D | Temperature environmental factor | Above 25 =1 Below 25=0 |
(54)
|
| dm | mRNA decay rate (1/s) | 0.00203 |
(54)
|
| β₁ | Antitoxin translation rate (1/s) | 0.139 |
(54)
|
| αTH | Binding of antitoxin and toxin through the high affinity site (T) (1/s) | 806.86 |
(54)
|
| γd | Unbinding of antitoxin and toxin through the high affinity site (T) (1/s) | 0.01773 |
(54)
|
| da | Antitoxin decay rate (1/s) | 0.212 |
(54)
|
| β₂ | Toxin translation rate (1/s) | 0.033 |
(54)
|
| dc | Decay rate due to cell cycle dilution (1/s) | 0.053 |
(54)
|
| F | Toxin release fraction (1/s) | 0.2 |
(54)
|
This equation models the mRNA transcribed from our switch. The mRNA level increases through transcription, represented by pF × D, which depends on the phosphorylation state of the promoter. When the promoter is phosphorylated, the transcription rate is ρu = 0.1333, while in the unphosphorylated state, the transcription rate is ρB = 0.
Temperature is another environmental factor included in the model. It is represented as a binary condition:
Above 25 °C = 1
Below 25 °C = 0
This equation models the anti-toxin (PemI). Its concentration increases through the translation rate (β₁) and the antitoxin unbinding rate (γd). It decreases due to the degradation rate (da) and the antitoxin binding rate (αTH).
This equation models toxin (PemK):
-It increases due to translation rate (β2), AT unbinding rate (γd), and toxin release fraction (F).
-It decreases due to degradation rate (dc) and AT binding rate (αTH).
This equation models the toxin antitoxin complex (PemIK):
-It increases due to AT binding rate (αTH), AT unbinding rate (γd), and toxin release fraction.
-It decreases due to degradation rate (dc), AT unbinding rate (γd), and toxin release fraction.
This figure shows PemIK kill switch component changes over time equal to 10 hours and the deactivation of the circuit at time (5h) that leads to decreased complex and antitoxin concentration and toxin accumulation.
This figure shows mRNA change over time before and after deactivation of the circuit.
The mRNA is transcribed from our circuit in the active condition that occurs only in the presence of the bacteria inside the lung environment. In other conditions the circuit will be inactive, and there will be no new mRNA transcription, and degradation will take the upper hand.
This figure shows PemI (Antitoxin) change over time before and after deactivation of the circuit and shows that steady state before deactivation occurs after 4h of activation and decreases after deactivation.
The PemI mRNA is translated in the active condition that occurs only in the presence of the bacteria inside the lung environment.In other conditions the circuit will be inactive, which leads to a decrease in mRNA concentration that PemI concentration by decreasing it.
This figure shows PemIK (the toxin antitoxin complex) change over time before and after deactivation. The complex shows the toxin antitoxin complex reaches zero after 5 h of deactivation.
This figure shows free PemK (toxin) concentration change over time.
Free toxin remains nearly zero under active conditions, and even after deactivation, any remnants quickly disappear due to the very high binding rate compared to the toxin translation rate.
Toxins begin to accumulate about 1.6 hours after deactivation, reach their maximum concentration at 2.29 hours, and then decline to zero by around 5.1 hours after deactivation.
We performed comprehensive sensitivity analysis across the main kinetic parameters, in order to understand how parameter variation can affect the robustness of our system and identify critical control points.
This figure shows Decay rate sensitivity – maximum toxin (kill efficiency).
The heatmap shows how PemI decay rate (da) and PemK decay rate (dc) affect the level of maximum free toxin, that determines killing efficiency of bacteria in other conditions in which the bacteria is not within the lung environment across parameter combination of PemI decay rate (da) and PemK decay rate (dc).
High toxin accumulation occurs when PemI decay rate (da) is high and PemK decay rate (dc) is low.
Sharp transitions indicate switch-like behavior, not a gradual one.
The white dashed lines position the system near the critical transition zone.
Our switch system show high kill efficiency as da = 4 * dc.
This figure shows translation rate sensitivity: Total protein ratio (A/T)
This analysis shows how PemI translation rate (B1) and PemK translation rate (B2) affect the critical antitoxin/toxin balance, the map shows the parameter space of (B1) vs (B2).
The system shows clear survival by (green) vs death (red) regions.
The system shows balanced synthesis conditions.
Small parameter changes can shift the survival system dramatically.
This figure shows translation rate sensitivity impact on free toxin at switch (kill efficiency)
This plot shows parameter combinations of PemI translation rate (B1) and PemK translation rate (B2) effect on maximum toxin indicating killing efficiency.
The narrow blue region represents a specific parameter space, where the biological switch can be activated or triggered.
The Sharp boundaries indicate bistable switch behavior.
The low translation rates of PemI and PemK are sufficient for switch function and shows high killing efficiency.
The sensitivity analysis reveals:
The requirements for parameter precision for reliable switch operation.
The critical parameters that we must control tightly (antitoxin degradation rate and translation balance).
A robust operation windows, where small variations don’t affect the function.
The failure modes, where minimal parameter drift can compromise safety.
We will fuse our PemIK control promoter with a GFP reporter gene to measure GFP expression under all 4 conditions. What we expect is GFP expression in lung conditions and absence in other conditions.
We will culture our LB bacteria under the four conditions and monitor CFU/time. What we expect is normal growth in lung conditions and bacteria death in other conditions.
To confirm the temperature cutoff, we plan to test the temperature response curve through a GFP reporter gene.
Our sensitivity analysis has identified critical parameters, that requires precise measurements :
PemI/pemK translation rate ratio (the most sensitive one to system behavior).
PemI & PemK degradation rate (It controls switch timing).
Complex binding and unbinding rate (affects toxin accumulation).
The priority experiments should focus on measuring these parameters with high accuracy.
1-It helped us to validate our circuit and gave confidence in biosafety before wet-lab validation.
2-Design of biosensor: It confirmed that phoB (phosphate) and RNA 2U (temperature) are suitable for our project, as they minimize survival of LB in other conditions where bacteria are outside the lung environment.
3-Parameter prioritization: Our sensitivity analysis model showed the priorities of the parameters, meaning experiments must measure them, such as PemI and PemK translation and transcription rates.
Guided experiments:It showed the future experimental validation plan we need to follow to increase confidence in our safety system.
This PemIK switch model demonstrates how combining a phosphate sensor and a temperature sensor is able to create a powerful safety device that ensures bacteria survive inside the lung environment only.
In this model with ODEs, we validated that:
The balance between toxin and antitoxin depends on promoter activity.
Free toxin accumulation occurs in conditions where bacteria are not inside the lung environment.
Bacterial death or growth is linked to free toxin accumulation.
Our PemIK safety system shows powerful switch-like behavior with clear safety margins.
The critical parameter combinations that we must be careful with if we deal with its value range, in order to prevent safety failure.
Optimal safe functional parameters zone for reliable lung-specific function.
This integrated models assessment reflects the excellence of our PULMORA modeling system in addressing the four iGEM criteria.
PULMORA model is the first integrated model to cover probiotic manufacturing all the way to clinical outcomes.
Scope: This model includes Three main chapters spanning three biological scales.
Technical Rigor: We combined deterministic predictions with Monte Carlo analysis to capture parameter uncertainty and improve confidence.
Clinical Relevance: It calculates bacterial viability from the lab bench, through freeze-drying, and finally to deposition in the bronchoalveolar region.
We validated every step in the project and made huge decisions that directed our project.
Freeze-drying as the most sensitive step in manufacturing.
The pH & H₂O₂ double-condition system, which prevents off-target activation and reduces basal leakage.
The rate of endosomal escape, which limits RNA therapeutic delivery.
The importance of precise control over bacterial proliferation to maintain balance.
The PemIK switch, as our safety device.
We integrated data from literature references, other iGEM teams like Technion 2019 and consultations with EIPICO, a pharmaceutical company.
Calibration of pH and H₂O₂ sensors in BALF under lung conditions.
RNA therapeutic kinetics in primary cells.
Long-term bacterial persistence inside capsules during storage.
Calibration of our bacterial strain’s viability and resistance in the lung.
Freeze-drying conditions specific to our strain.
Yes, because it provides:
Methodological contributions: Cascade modeling that clearly shows how upstream model outputs affect downstream predictions.
Clinical translation: We showed how a synthetic biology design can directly connect to a real therapeutic benefit—by targeting and breaking down the root cause of asthma, TSLP.
Honest assessment: We openly recognized where our model has limits and clearly pointed out the validation experiments still needed to move forward.
Reproducibility:
All parameters are sourced and referenced; uncertain ones are strengthened through Monte Carlo simulation.
MATLAB and colab codes are openly available for others in the community.
A clear experimental validation plan for future work.
Alignment with pharmaceutical development processes, bridging synthetic biology with real-world standards.
CO-BERA use goes beyond asthma as it orean be repurposed to target cancer drivers, viral infections, autoimmune pathways, or metabolic diseases. Future iGEM teams can adopt our composite design to load their own siRNA sequences into CO-BERA, achieving precision. We provide the community with a versatile RNA-based therapeutic chassis that can serve as a general platform for synthetic biology applications.
Design Optimization: We have informed 12 critical decisions throughout development that directly influenced our design.
Validation: We ensured validation of all our project systems and steps.
Experimental Prioritization: We have defined validation experiments and arranged them by priority, based on impact and feasibility.
Experimental Validation: Around 30% of the parameters still require direct measurement specific to our conditions and bacterial strain.
Patient Stratification: Not all individuals are the same in pharmacokinetics and response, which needs to be addressed.
High Priority: Calibration of pH and H₂O₂ sensors in BALF under lung conditions.
Medium Priority: Lung persistence studies of our bacterial strain inside the lung.
Long-Term: Patient stratification models across larger populations, with correlation to clinical biomarkers such as IL-5.
Our PULMORA model provides a solid quantitative foundation that carries development from concept toward clinical reality. We highlighted the experimental validation needed, yet it still needs to be completed.
Struhk, A., Smelt, J. P., & Brul, S. (2000). Effects of cryoprotective agents on survival of Listeria monocytogenes, Lactobacillus plantarum and Lactobacillus rhamnosus strains after freeze-drying and storage. CryoLetters, 21(6), 347–356.
Madigan, M. T., Martinko, J. M., Stahl, D. A., & Clark, D. P. (2012). Brock biology of microorganisms (13th ed.). Benjamin Cummings.
Pikal, M. J. (2002). Freeze-drying of proteins: Process, formulation, and stability. In Formulation and delivery of proteins and peptides (pp. 120–121). American Chemical Society.
Carvalho, A. S., Silva, J., Ho, P., Teixeira, P., Malcata, F. X., & Gibbs, P. (2004). Relevant factors for the preparation of freeze-dried lactic acid bacteria. International Dairy Journal, 14(10), 835-847.
Jennings, T. A. (2002). Lyophilization: introduction and basic principles. CRC press.
Doran, P. M. (2013). Bioprocess engineering principles (2nd ed.). Academic Press.
Dalby, R. N., Byron, P. R., Hindle, M., Peart, J., Farr, S. J., & Wauthoz, N. (2000). Development of dry powder inhalers. Pharmaceutical Technology, 24(10), 116–128.
Thakur, A., Xu, Y., Cano-Garcia, G., Feng, S., Rose, F., Gerde, P., Andersen, P., Christensen, D., & Foged, C. (2022). Optimizing the design and dosing of dry powder inhaler formulations of the cationic liposome adjuvant CAF®01 for pulmonary immunization. Frontiers in Drug Delivery, 2, Article 973599.
Hickey, A. J., & Mansour, H. M. (Eds.). (2019). Inhalation Aerosols: Physical and Biological Basis for Therapy. CRC Press.
Pilcer, G., & Amighi, K. (2010). Formulation strategy and use of excipients in pulmonary drug delivery. International Journal of Pharmaceutics, 392(1–2), 1–19
Cui, J., Li, W., Xu, T., & Sun, Y. (2020). Viability of probiotics in spray-dried and freeze-dried formulations containing trehalose. Drying Technology, 38(12), 1594–1603.
Cheow, W. S., & Hadinoto, K. (2013). Enhancing encapsulation and aerosolization of viable bacterial cells. Journal of the Royal Society Interface, 10(84), 20130143.
Usmani, O. S., Biddiscombe, M. F., & Barnes, P. J. (2003). Regional lung deposition and bronchodilator response as a function of β2-agonist particle size. American Journal of Respiratory and Critical Care Medicine, 168(8), 914–920.
ICRP Publication 66. (1994). Human Respiratory Tract Model for Radiological Protection. Annals of the ICRP, 24(1-3).
Wanner, A., Salathé, M., & O'Riordan, T. G. (1996). Mucociliary clearance in the airways. American Journal of Respiratory and Critical Care Medicine, 154(6), 1868-1902.
Thomas, P. S., Herbert, H., & Hill, D. J. (1998). Effect of an inhaled corticosteroid on mucociliary clearance and airway inflammation in mild asthma. Thorax, 53(10), 858-862.
Forsyth, A., Weeks, C., & Rich, G. (2012). Lactobacillus plantarum NCIMB8826 is killed by alveolar macrophages from rats. Clinical and Experimental Immunology, 168(2), 226–232.
Alexis, N. E., Soukup, J., Nierkens, S., Becker, S., & Cascio, W. E. (2006). Macrophage secretion of proinflammatory cytokines following exposure to urban air particulates. Journal of Toxicology and Environmental Health, Part A, 69(7), 587–601.
Diamond, G., Beckloff, N., Ryan, L. K., & Kolls, J. K. (2008). Host defense peptides in the lung: an untapped arsenal. Current pharmaceutical design, 14(13), 1268-1279.
Rojas-Rejón, Ó. A., Gonzalez-Figueredo, C., Quintero-Covarrubias, A. R., & Saldaña-Jáuregui, A. (2024). Growth of Lactiplantibacillus plantarum BG112 in batch and continuous culture with Camellia sinensis as prebiotic. Fermentation, 10(9), Article 487.
Letizia, F., Albanese, G., Testa, B., Vergalito, F., Bagnoli, D., Di Martino, C., Carillo, P., Verrillo, L., Succi, M., Sorrentino, E., Coppola, R., Tremonte, P., Lombardi, S. J., Di Marco, R., & Iorizzo, M. (2022). In vitro assessment of bio-functional properties from Lactiplantibacillus plantarum strains. Current Issues in Molecular Biology, 44(5), 2321–2334.
Park,M. Choi, S. & Lee, Y. 2020. Evaluation of Lactobacillus plantarum K8 in a murine model of respiratory infection probiotic colonization and host immune regulation. Frontiers in Microbiology, 11, 590.
Glieca, S., Quarta, E., Bottari, B., Bancalari, E., Monica, S., Scaltriti, E., Tambassi, M., Flammini, L., Bertoni, S., Bianchera, A., Fainardi, V., Esposito, S., Pisi, G., Bettini, R., Sonvico, F., & Buttini, F. (2024). Development of inhalation powders containing lactic acid bacteria with antimicrobial activity against Pseudomonas aeruginosa. International Journal of Antimicrobial Agents, 63, 107001.
Wusqy, N. K., & Surono, I. S. (2021). Shelf life of microencapsulated and free cells of Lactobacillus plantarum IS-10506 at different temperatures. IOP Conference Series: Earth and Environmental Science, 794(1), 012147.
Kaymak Ertekin, F., et al. (2024). Improvement of freeze-dried survival of Lactobacillus plantarum BG24. Applied Microbiology & Biotechnology, 108(10), 3801–3810.
Santivarangkna, C., Kulozik, U., & Först, P. (2008). Inactivation mechanisms of lactic acid starter cultures preserved by drying processes. Journal of Applied Microbiology, 105(1), 1–13.
Ananta, E., Volkert, M., & Knorr, D. (2005). Cellular injuries and storage stability of spray-dried Lactobacillus rhamnosus GG. International Journal of Food Science & Technology, 40(6), 599–607.
Silva, J., Corcoran, B. M., Ross, R. P., Fitzgerald, G. F., & Stanton, C. (2002). Comparative survival of probiotic lactobacilli spray-dried in the presence of prebiotic substances. Journal of Applied Microbiology, 92(6), 1024–1032.
ICH Q1A(R2) (2003). Stability Testing of New Drug Substances and Products. International Conference on Harmonisation.
Sulabo, A. S. L., Villasanta, M. E. L., Hermo, K. G., Lascano, R. A., Collado, L. S., & Babaran, G. M. O. (2020). Storage stability of freeze-dried Lactobacillus plantarum S20 starter culture as affected by various formulations of drying medium, and its fermentation characteristics in mung bean (Vigna radiata L.) slurry. Food Research, 4(4), 964–975.
Flick, G. J., & Kuenen, J. G. (1999). Activation energy of microbial inactivation for Lactobacillus species in freeze-dried probiotic formulations. International Journal of Food Microbiology, 47(1–2), 67–74.
IUPAC (1997). Quantities, Units and Symbols in Physical Chemistry (2nd ed.). International Union of Pure and Applied Chemistry.
Carstensen, J. T., & Rhodes, C. T. (2000). Drug stability: Principles and practices (3rd ed.). Marcel Dekker.
Rowland, M., & Tozer, T. N. (2011). Clinical Pharmacokinetics and Pharmacodynamics: Concepts and Applications (4th ed.).
Jensen, P. R., & Hammer, K. (1998). The sequence of spacers between the consensus half-sites modulates the strength of prokaryotic promoters. Applied and Environmental Microbiology, 64(1), 82–87.
Driessen, A. J. M., van der Does, C., & Poolman, B. (2001). Two acid-inducible promoters from Lactococcus lactis require the cis-acting ACiD-box and the transcription regulator RcfB. Molecular Microbiology, 49(3), 740–752.
Even, S., Pellegrini, O., Zig, L., Labas, V., Vinh, J., Bréchemmier-Baey, D., & Putzer, H. (2005). Ribonucleotide reductase is regulated by the pH-dependent activity of the Lactococcus lactis nrd promoter. Journal of Bacteriology, 187(16), 5837–5845.
iGEM. (2019). Team: SZPT-CHINA – Model.
iGEM. (2015). Team: BABS_UNSW_Australia – Project: pHlow.
Kelly, J. R., Rubin, A. J., Davis, J. H., Ajo-Franklin, C. M., Cumbers, J., Czar, M. J., de Mora, K., Glieberman, A. L., Monie, D. D., & Endy, D. (2009). Measuring the activity of BioBrick promoters using an in vivo reference standard. Journal of Biological Engineering, 3(4), 4.
Densmore, D., & Endy, D. (2013). Promoter element arising from the fusion of standard BioBrick parts. ACS Synthetic Biology, 2(2), 111–120.
Canton, B., Labno, A., & Endy, D. (2008). Refinement and standardization of synthetic biological parts and devices. Nature Biotechnology, 26(7), 787–793.
iGEM. (2019). Team: Technion-Israel – Model.
Ackers, G. K., Johnson, A. D., & Shea, M. A. (1982). Quantitative model for gene regulation by lambda phage repressor. Proceedings of the National Academy of Sciences, 79(4), 1129–1133.
Sahay, G., Querbes, W., Alabi, C., Eltoukhy, A., Sarkar, S., Zurenko, C., Karagiannis, E., Love, K., Chen, D., Zoncu, R., Buganim, Y., Schroeder, A., Langer, R., & Anderson, D. G. (2013). Efficiency of siRNA delivery by lipid nanoparticles is limited by endocytic recycling. Nature Biotechnology, 31(7), 653–658.
Gupta, A., & Mandal, S. (2020). Improving the endosomal escape of biologics. Advanced Drug Delivery Reviews, 160, 2–12.
Zamore, P. D., Tuschl, T., Sharp, P. A., & Bartel, D. P. (2000). RNAi: Double-stranded RNA directs the ATP-dependent cleavage of mRNA at 21 to 23 nucleotide intervals. Cell, 101(1), 25–33.
Sajid, M. I., Moazzam, M., & Cho, K.-S. (2020). Stability and delivery of siRNA therapeutics. Pharmaceuticals, 13(11), 374.
Zhao, X., Wei, Y., Bu, Y., Ren, X., & Dong, Z. (2025). Review on bacterial outer membrane vesicles: Structure, vesicle formation, separation and biotechnological applications. Microbial Cell Factories, 24, 27.
Zanders, E. D. (2017). Delivery of RNA therapeutics: The great endosomal escape! Molecular Therapy, 25(8), 1719–1720.
Haley, B., & Zamore, P. D. (2004). Kinetic analysis of the RNAi enzyme complex. Nature Structural & Molecular Biology, 11(7), 599–606.
Larsson, E., Sander, C., & Marks, D. (2010). mRNA turnover rate limits siRNA and microRNA efficacy. Molecular Systems Biology, 6, 433.
Maldonado-Barragán, A., Ruiz-Barba, J. L., & Jiménez-Díaz, R. (2009). Knockout of three-component regulatory systems reveals that the apparently constitutive plantaricin-production phenotype shown by Lactobacillus plantarum on solid medium is regulated via quorum sensing. International Journal of Food Microbiology, 130(1), 35-42.
Gelens, L., Hill, L., Vandervelde, A., Danckaert, J., & Loris, R. (2013). A general model for toxin- antitoxin module dynamics can explain persister cell formation in Escherichia coli. PLoS Computational Biology, 9(8), e1003190.
Sportès, C., Hakim, F. T., Mitchell, D., Wang, H. L., Jantz, K., Cleveland, J. L., & Gress, R. E. (2008). Administration of recombinant human interleukin-7 to nonhuman primates: pharmacokinetics and biologic effects. The Journal of Clinical Investigation, 118(4),.
Bartlett, D. W., & Davis, M. E. (2006). Insights into the kinetics of siRNA-mediated gene silencing from live-cell imaging. Biotechnology and Bioengineering, 93(6), 1210–1217.
Sportès, C., Hakim, F. T., Mitchell, D., Wang, H. L., Jantz, K., Cleveland, J. L., & Gress, R. E. (2008). Administration of recombinant human interleukin-7 to nonhuman primates: pharmacokinetics and biologic effects. The Journal of Clinical Investigation, 118(4),.
To ensure that our Co-BERA scaffold could be efficiently delivered, accurately identify, and effectively bind to its target TSLP mRNA, we needed to model its interactions with various molecular partners. This process helped us select the most suitable candidate capable of achieving the desired TSLP mRNA silencing.
We employed a combination of molecular docking, RNA structural analysis, and molecular dynamics simulations to evaluate potential candidates based on their structural stability and ability to enhance the scaffold's silencing efficiency.
Each step of this process was thoroughly documented with screenshots, serving as a tutorial guide for future teams interested in performing molecular dynamics simulations. Additionally, we created a 3D inhaler model to visually represent and summarize the outcomes of our five main simulation models.
(2021). Highly accurate protein structure prediction with AlphaFold. Nature, 596(7873), 583-589.
(2020). AMBER 2020. University of California, San Francisco.
(2013). CHARMM-GUI micelle builder for pure/mixed micelle and protein/micelle complex systems. Journal of Chemical Information and Modeling, 53(8), 2171-2180.
(2007). Automated builder and database of protein/membrane complexes for molecular dynamics simulations. PLoS ONE, 2(9), e880.
(2009). CHARMM-GUI Membrane Builder for mixed bilayers and its application to yeast membranes. Biophysical Journal, 97(1), 50-58.
(2019). CHARMM-GUI Membrane Builder for complex biological membrane simulations with glycolipids and lipoglycans. Journal of Chemical Theory and Computation, 15(1), 775-786.
(2021). CHARMM-GUI Membrane Builder for lipid nanoparticles with ionizable cationic lipids and PEGylated lipids. Journal of Chemical Information and Modeling, 61(10), 5192-5202.
(2014). CHARMM-GUI Membrane Builder toward realistic biological membrane simulations. Journal of Computational Chemistry, 35(27), 1997-2004.
(2008). CHARMM-GUI: A web-based graphical user interface for CHARMM. Journal of Computational Chemistry, 29(11), 1859-1865.
(2016). CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. Journal of Chemical Theory and Computation, 12(1), 405-413.
(2020). CHARMM-GUI supports the Amber force fields. The Journal of Chemical Physics, 153(3), 035103.
(2024). Lactiplantibacillus plantarum WCFS1, complete sequence [Genome sequence NC_004567.2]. GenBank.
(n.d.). LSU ribosomal protein L7AE (rpl7AE) [Archaeoglobus fulgidus DSM 4304] [Protein sequence AAB90466.1]. GenBank.
(2015). GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX, 1-2, 19-25.
(1995). GROMACS: A message-passing parallel molecular dynamics implementation. Computer Physics Communications, 91(1-3), 43-56.
(2008). GROMACS 4: Algorithms for highly efficient, load-balanced, and scalable molecular simulation. Journal of Chemical Theory and Computation, 4(3), 435-447.
(2001). GROMACS 3.0: A package for molecular simulation and trajectory analysis. Journal of Molecular Modeling, 7(8), 306-317.
(2015). Tackling exascale software challenges in molecular dynamics simulations with GROMACS. In S. Markidis & E. Laure (Eds.), Solving software challenges for exascale (pp. 3-27). Springer.
(2005). GROMACS: Fast, flexible, and free. Journal of Computational Chemistry, 26(16), 1701-1718.
(2017). HDOCK: A web server for protein-protein and protein-DNA/RNA docking based on a hybrid strategy. Nucleic Acids Research, 45(W1), W365-W373.
(n.d.). Lipidomic characterization and localization of phospholipids in the human lung [Supplementary material]. Department of Pharmacology, University of Colorado Denver; Department of Medicine, National Jewish Health.
(2015). Effect of storage's temperature on phospholipids compositions and viability upon Lactobacillus plantarum CWB-B1419. International Journal of Science and Research, 4(9), 438-445.
(2015). MDTraj: A modern open library for the analysis of molecular dynamics trajectories. Biophysical Journal, 109(8), 1528-1532.
(2017). OpenMM 7: Rapid development of high performance algorithms for molecular dynamics. PLOS Computational Biology, 13(7), e1005659.
(n.d.). PyMOL command reference.
(2009). siDirect 2.0: Updated software for designing functional siRNA with reduced seed-dependent off-target effect. BMC Bioinformatics, 10, 392.
(2018). SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Research, 46(W1), W296-W303.
(n.d.-b). Protein argonaute-2 (Q9UKV8). UniProt Knowledgebase.
(n.d.-a). DUF4811 domain-containing protein [Lactiplantibacillus plantarum subsp. plantarum P-8] [Protein entry F9URU2]. UniProt Knowledgebase.
(2013). Crystal structure of a zinc enzyme, 4CDB. RCSB Protein Data Bank.
(2011). ViennaRNA Package 2.0. Algorithms for Molecular Biology, 6, 26.
Logic gates are a fundamental component of digital circuits that perform basic logical operations on one or more binary inputs to generate a single binary output. The output varies depending on the type of logic gate and the combination of inputs. The output and the input are in one of the two states:
Biological circuits can be used to regulate cellular response to environmental conditions, cellular development, or modify its function. Biological circuits can perform logical operations similar to electronic circuits. Therefore, we use the logic gates as biological circuits to mimic the function of electronic logic gates and predict the output based on different input combinations.
PhoB: It’s a constitutive promoter, which is inhibited when extracellular phosphate reaches 50 µM or more. Thus, limiting the expression of toxin and antitoxin.
PhoR: it is a histidine kinase sensor protein that is associated with the inner membrane. It works by dephosphorylating the PhoB, inhibiting its action, and inhibiting both toxin and antitoxin expression.
Thermosensor: It’s a non-coding heat-inducible RNA 2U that acts as a heat-inducible thermosensor and stops the expression of the following gene below 25°C.
In the low of PO₄ and at temperatures above 25°C: PhoB and the thermosensor act normally and express both toxin and antitoxin. Then, the antitoxin will neutralize the toxin, leaving the bacteria to live normally.
In high PO₄ and a temperature above 25°C: PhoR will constrain the action of PhoB, and the thermosensitive RNA will fold at the ribosomal binding site. As a result, it inhibits both toxin and antitoxin expression. Then the antitoxin will be degraded by Lon protease, which allows the toxin to kill the bacteria.
In high PO₄ only: The phoB will stop the expression of both toxin and antitoxin.
At temperatures below or equal to 25°C only: The thermosensor RNA will fold at the ribosomal binding site, inhibiting both toxin and antitoxin expression.
We consider the input to be phosphate.
Since the thermosensor is active at temperatures above 25°C, we considered setting our input to be above 25°C.
This figure illustrates the toxin and antitoxin in condition above 25°C
There are two inputs: the first is the output of the NOT gate, and the second is the output from the thermosensor. The inputs of the (AND) gate are the outputs of the previous gates. We need to express the toxin and antitoxin, so both inputs should be 1 (ON). At this point we have many scenarios:
PKatA promoter: is a Per repressor-regulated DNA sequence that is designed to function as a biosensor (which senses H₂O₂) in Gram-positive bacteria.
PKatA is employed to sense H₂O₂ by the Per repressor. So, when the amount of H₂O₂ is increased, the expression of LLO will increase.
This gate is responsible for the expression of LLO. We consider the input to be per repressor. Expression of LLO depends on activation of the pKatA promoter, which depends on the state of the Per repressor. The Per repressor inhibits pKatA expression activity in the low H₂O₂. The Per repressor becomes inactive in high H₂O₂, in this condition pKatA has a high expression activity. So, we have two scenarios:
LLO secreted from bacteria is inactive. Therefore, when membrane vesicles fuse with endosomes, which have acidic media (pH=5.5), LLO is activated and permits the bacteria to escape into the cytosol.
There are two inputs: the first is the output of the NOT gate, and the second is the output of the presence of acidity or not. We need to activate the LLO, so the only scenario is when both inputs are 1. Hence, from this gate there are 4 scenarios:
p170-CP25 promoter:it initiates transcription of the Lac repressor under acidic conditions (pH below 6.9).
p32 promoter: It drives constitutive transcription of Rep repressor.
PKatA promoter: In this circuit, PKatA is a double regulated DNA sequence that is designed to function as a biosensor (which senses H₂O₂) in Gram-positive bacteria, it also has an operator in its sequence that binds to Rep repressor if it is present.
Lac Repressor: Its expression leads to inhibiting the p32 promoter and stops the expression of Rep Repressor.
Rep Repressor: when it is expressed, it stops pKatA from expressing the CO-BERA.
Per Repressor: it is able to sense low levels of hydrogen peroxide (H₂O₂). Per repressor senses H₂O₂ by metal-catalyzed oxidation.
For CO-BERA expression, there are two stages. The first stage is to stop the expression of the Rep repressor, which inhibits the pKatA promoter under the effect of acidic pH. The second is to inactivate the Per repressor under the effect of H₂O₂.
This circuit consists of four NOT gates and one AND gate to express CO-BERA. NOT gates (No. 1 & 2) were formed to stop the expression of the repressor. On the other hand, NOT gates (No. 3 & 4) were formed to inactivate the Per repressor. As a result, both of these outputs form the inputs of the AND gate to activate the pKatA promoter. They are explained as follows:
In NOT gate one, the input is the Lac repressor, which will be expressed when pH becomes below 6.9. So, the output of this gate is the inactivation of the p32 promoter, which leads to inhibiting the expression of the Rep repressor. Meaning that the input is 1, and the output is 0.
In NOT gate two, the input for it shows no expression of the repressor. Thus, the output is that pKatA is available to initiate the transcription of CO-BERA. As a result, the input is 1, and the output will be 0. Meaning that the pKatA promoter can do its function and that it can’t happen without the presence of a high amount of H₂O₂.
This gate is complementary to the previous gates. In addition, the input is H₂O₂, which acts on the Per repressor:
In NOT gate three, the input is the H₂O₂ and the output will be Per repressor expression. In high H₂O₂ which is expressed in asthma condition the input will be 0 and the output will be 1, that means Per repressor will be expressed. In normal condition there is low H₂O₂ so the input will be 0 and the output will be 1, that means the Per repressor will not be expressed.
In NOT gate four, the input is the Per repressor expression and the output will be input for AND gate. In the presence of Per repressor the input will be 0 and the output will be 1. In the absence of Per repressor so the input will be 0 and the output will be 1.
We have two inputs. One from pH and the second from H₂O₂. They are demonstrated in four scenarios: