Path: blob/main/latex-templates/templates/chemical-engineering/reaction_engineering.tex
51 views
unlisted
\documentclass[11pt,a4paper]{article}1\usepackage[utf8]{inputenc}2\usepackage[T1]{fontenc}3\usepackage{amsmath,amssymb}4\usepackage{graphicx}5\usepackage{booktabs}6\usepackage{siunitx}7\usepackage{geometry}8\geometry{margin=1in}9\usepackage{pythontex}10\usepackage{hyperref}11\usepackage{float}12\usepackage{cite}1314\title{Reaction Engineering\\CSTR and PFR Design}15\author{Chemical Engineering Research Group}16\date{\today}1718\begin{document}19\maketitle2021\begin{abstract}22This report presents comprehensive computational analysis of chemical reactor design and kinetics, focusing on batch reactors, continuous stirred tank reactors (CSTRs), and plug flow reactors (PFRs). We investigate first-order and second-order reaction kinetics, residence time distributions, temperature effects via Arrhenius behavior, and reactor performance comparisons using Levenspiel plots. The analysis provides design equations, numerical solutions for non-isothermal operation, and optimization strategies for industrial reactor systems.23\end{abstract}2425\section{Introduction}2627Chemical reactor design is fundamental to the chemical process industry, determining the size, configuration, and operating conditions required to achieve desired conversion and selectivity \cite{Fogler2016, Levenspiel1999}. The choice between batch, CSTR, and PFR configurations depends on reaction kinetics, heat transfer requirements, and production scale \cite{Smith2005}.2829This computational study examines reactor performance across different configurations, investigating the effects of reaction order, temperature, and residence time on conversion. We employ Python-based numerical methods to solve design equations and generate Levenspiel plots for reactor comparison \cite{Scott2006}.3031\begin{pycode}3233import numpy as np34import matplotlib.pyplot as plt35from scipy.integrate import odeint, solve_ivp36from scipy.optimize import fsolve, brentq37plt.rcParams['text.usetex'] = True38plt.rcParams['font.family'] = 'serif'3940# Global constants41R = 8.314 # J/(mol·K) - Universal gas constant4243\end{pycode}4445\section{Batch Reactor Kinetics}4647For a constant-volume batch reactor, the design equation for species A is:48\begin{equation}49\frac{dC_A}{dt} = -r_A50\end{equation}5152For a first-order irreversible reaction $A \rightarrow P$ with rate constant $k$:53\begin{equation}54-r_A = k C_A \quad \Rightarrow \quad C_A(t) = C_{A0} e^{-kt}55\end{equation}5657For a second-order reaction $A \rightarrow P$:58\begin{equation}59-r_A = k C_A^2 \quad \Rightarrow \quad C_A(t) = \frac{C_{A0}}{1 + k C_{A0} t}60\end{equation}6162The conversion $X_A$ is defined as: $X_A = 1 - C_A/C_{A0}$6364\begin{pycode}65# Batch reactor kinetics comparison66k1 = 0.1 # rate constant for first-order (1/min)67k2 = 0.15 # rate constant for second-order (L/(mol·min))68C_A0 = 1.0 # initial concentration (mol/L)69t = np.linspace(0, 60, 500)7071# First-order kinetics72C_A_first = C_A0 * np.exp(-k1 * t)73X_A_first = 1 - C_A_first / C_A07475# Second-order kinetics76C_A_second = C_A0 / (1 + k2 * C_A0 * t)77X_A_second = 1 - C_A_second / C_A07879# Store key results80t_90_first = -np.log(0.1) / k181t_90_second = 9 / (k2 * C_A0)8283fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))8485# Concentration profiles86ax1.plot(t, C_A_first, 'b-', linewidth=2, label='First-order')87ax1.plot(t, C_A_second, 'r--', linewidth=2, label='Second-order')88ax1.axhline(y=0.1*C_A0, color='gray', linestyle=':', alpha=0.5)89ax1.set_xlabel('Time (min)', fontsize=11)90ax1.set_ylabel('$C_A$ (mol/L)', fontsize=11)91ax1.set_title('Batch Reactor Concentration Decay', fontsize=12)92ax1.legend(fontsize=10)93ax1.grid(True, alpha=0.3)9495# Conversion profiles96ax2.plot(t, X_A_first, 'b-', linewidth=2, label='First-order')97ax2.plot(t, X_A_second, 'r--', linewidth=2, label='Second-order')98ax2.axhline(y=0.9, color='gray', linestyle=':', alpha=0.5, label='90\\% conversion')99ax2.set_xlabel('Time (min)', fontsize=11)100ax2.set_ylabel('Conversion $X_A$', fontsize=11)101ax2.set_title('Batch Reactor Conversion', fontsize=12)102ax2.legend(fontsize=10)103ax2.grid(True, alpha=0.3)104105plt.tight_layout()106plt.savefig('reaction_engineering_plot1.pdf', dpi=150, bbox_inches='tight')107plt.close()108\end{pycode}109110\begin{figure}[H]111\centering112\includegraphics[width=0.95\textwidth]{reaction_engineering_plot1.pdf}113\caption{Batch reactor performance for first-order and second-order reactions. First-order kinetics exhibit exponential decay, while second-order reactions show hyperbolic behavior. For 90\% conversion with $C_{A0} = \py{C_A0}$ mol/L, first-order requires $\py{f'{t_90_first:.1f}'}$ min while second-order requires $\py{f'{t_90_second:.1f}'}$ min, demonstrating the impact of reaction order on batch time.}114\end{figure}115116\section{CSTR Design}117118For a continuous stirred tank reactor (CSTR) at steady state, the design equation is:119\begin{equation}120\tau = \frac{V}{Q} = \frac{C_{A0} - C_A}{-r_A} = \frac{C_{A0} X_A}{-r_A}121\end{equation}122123For a first-order reaction in a single CSTR:124\begin{equation}125\tau = \frac{X_A}{k(1 - X_A)}126\end{equation}127128For $N$ equal-volume CSTRs in series, each with volume $V_i = V/N$:129\begin{equation}130X_{A,N} = 1 - \frac{1}{(1 + k\tau/N)^N}131\end{equation}132133\begin{pycode}134# CSTR design - single and multiple reactors in series135k_cstr = 0.1 # rate constant (1/min)136X_target = 0.90 # target conversion137138# Single CSTR residence time for target conversion139tau_single = X_target / (k_cstr * (1 - X_target))140141# Multiple CSTRs in series142N_reactors = np.array([1, 2, 3, 4, 5, 10])143X_A_range = np.linspace(0.01, 0.99, 200)144145fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))146147# Conversion vs residence time for different numbers of CSTRs148tau_values = np.linspace(0, 50, 300)149for N in [1, 2, 3, 5]:150if N == 1:151X_vals = (k_cstr * tau_values) / (1 + k_cstr * tau_values)152label_str = 'Single CSTR'153else:154X_vals = 1 - 1 / (1 + k_cstr * tau_values / N)**N155label_str = f'{N} CSTRs in series'156ax1.plot(tau_values, X_vals, linewidth=2, label=label_str)157158ax1.axhline(y=0.9, color='gray', linestyle=':', alpha=0.5)159ax1.axvline(x=tau_single, color='gray', linestyle=':', alpha=0.5)160ax1.set_xlabel('Total Residence Time $\\tau$ (min)', fontsize=11)161ax1.set_ylabel('Conversion $X_A$', fontsize=11)162ax1.set_title('CSTR Performance: Effect of Staging', fontsize=12)163ax1.legend(fontsize=9)164ax1.grid(True, alpha=0.3)165ax1.set_xlim(0, 50)166167# Volume requirement for different staging configurations168total_volumes = []169for N in N_reactors:170if N == 1:171tau_req = X_target / (k_cstr * (1 - X_target))172else:173# Solve for tau: X_A = 1 - 1/(1 + k*tau/N)^N174def equation(tau):175return 1 - 1/(1 + k_cstr*tau/N)**N - X_target176tau_req = brentq(equation, 0, 100)177total_volumes.append(tau_req) # Assuming Q=1 so V=tau178179ax2.plot(N_reactors, total_volumes, 'o-', linewidth=2, markersize=8, color='steelblue')180ax2.axhline(y=tau_single, color='gray', linestyle='--', alpha=0.5, label='Single CSTR')181ax2.set_xlabel('Number of CSTRs in Series', fontsize=11)182ax2.set_ylabel('Total Volume Required (L)', fontsize=11)183ax2.set_title(f'Volume Optimization for $X_A = {X_target}$', fontsize=12)184ax2.legend(fontsize=10)185ax2.grid(True, alpha=0.3)186187plt.tight_layout()188plt.savefig('reaction_engineering_plot2.pdf', dpi=150, bbox_inches='tight')189plt.close()190191# Store results192V_single = total_volumes[0]193V_three = total_volumes[2]194reduction_pct = (V_single - V_three) / V_single * 100195\end{pycode}196197\begin{figure}[H]198\centering199\includegraphics[width=0.95\textwidth]{reaction_engineering_plot2.pdf}200\caption{CSTR staging analysis showing the benefit of multiple reactors in series. For first-order kinetics with $k = \py{k_cstr}$ min$^{-1}$, achieving $X_A = \py{X_target}$ requires $\tau = \py{f'{tau_single:.1f}'}$ min for a single CSTR. Using three CSTRs in series reduces total volume by $\py{f'{reduction_pct:.1f}'}$\%, approaching PFR performance while maintaining the advantages of perfect mixing within each stage.}201\end{figure}202203\section{Plug Flow Reactor (PFR) Design}204205For a PFR, the differential design equation is:206\begin{equation}207\frac{dX_A}{dV} = \frac{-r_A}{F_{A0}}208\end{equation}209210For isothermal first-order kinetics, this integrates to:211\begin{equation}212V = \frac{F_{A0}}{k} \ln\left(\frac{1}{1-X_A}\right) = \frac{Q C_{A0}}{k} \ln\left(\frac{1}{1-X_A}\right)213\end{equation}214215Comparing residence times: $\tau_{PFR} = -\frac{1}{k}\ln(1-X_A)$ vs. $\tau_{CSTR} = \frac{X_A}{k(1-X_A)}$216217For non-isothermal operation, we must solve coupled ODEs numerically:218\begin{align}219\frac{dX_A}{dV} &= \frac{k(T) C_{A0} (1-X_A)}{F_{A0}} \\220\frac{dT}{dV} &= \frac{(-\Delta H_R) k(T) C_{A0} (1-X_A) - UA(T - T_c)}{\rho C_p Q}221\end{align}222223\begin{pycode}224# PFR design with non-isothermal operation225k_pfr = 0.1 # rate constant at reference T (1/min)226C_A0_pfr = 1.0 # inlet concentration (mol/L)227Q_pfr = 10.0 # volumetric flow rate (L/min)228F_A0 = Q_pfr * C_A0_pfr # molar flow rate (mol/min)229230# Isothermal PFR231X_A_pfr = np.linspace(0.01, 0.95, 100)232V_pfr = (F_A0 / k_pfr) * np.log(1 / (1 - X_A_pfr))233tau_pfr = V_pfr / Q_pfr234235# Compare with CSTR236tau_cstr_comp = X_A_pfr / (k_pfr * (1 - X_A_pfr))237V_cstr_comp = tau_cstr_comp * Q_pfr238239# Non-isothermal PFR with exothermic reaction240def pfr_nonisothermal(V, y):241X_A, T = y242243# Arrhenius temperature dependence244A = 1e10 # pre-exponential factor (1/min)245Ea = 75000 # activation energy (J/mol)246k_T = A * np.exp(-Ea / (R * T))247248# Energy balance parameters249DH_R = -50000 # heat of reaction (J/mol) - exothermic250rho = 1000 # density (kg/m³)251Cp = 4180 # heat capacity (J/(kg·K))252U = 500 # heat transfer coefficient (W/(m²·K))253A_heat = 0.5 # heat transfer area per volume (m²/L)254T_c = 298 # coolant temperature (K)255256dX_dV = (k_T * C_A0_pfr * (1 - X_A)) / F_A0257dT_dV = ((-DH_R) * k_T * C_A0_pfr * (1 - X_A) - U * A_heat * (T - T_c)) / (rho * Cp * Q_pfr / 1000)258259return [dX_dV, dT_dV]260261# Initial conditions262T_inlet = 320 # inlet temperature (K)263V_range = np.linspace(0, 50, 500)264265sol = odeint(pfr_nonisothermal, [0, T_inlet], V_range)266X_A_noniso = sol[:, 0]267T_profile = sol[:, 1]268269fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))270271# Isothermal reactor comparison272ax1.plot(tau_pfr, X_A_pfr, 'b-', linewidth=2, label='PFR')273ax1.plot(tau_cstr_comp, X_A_pfr, 'r--', linewidth=2, label='CSTR')274ax1.axhline(y=0.9, color='gray', linestyle=':', alpha=0.5)275ax1.set_xlabel('Residence Time $\\tau$ (min)', fontsize=11)276ax1.set_ylabel('Conversion $X_A$', fontsize=11)277ax1.set_title('PFR vs CSTR: Isothermal First-Order', fontsize=12)278ax1.legend(fontsize=10)279ax1.grid(True, alpha=0.3)280ax1.set_xlim(0, 50)281282# Non-isothermal PFR283ax2_temp = ax2.twinx()284ax2.plot(V_range, X_A_noniso, 'b-', linewidth=2, label='Conversion')285ax2_temp.plot(V_range, T_profile, 'r--', linewidth=2, label='Temperature')286ax2.set_xlabel('Reactor Volume (L)', fontsize=11)287ax2.set_ylabel('Conversion $X_A$', fontsize=11, color='b')288ax2_temp.set_ylabel('Temperature (K)', fontsize=11, color='r')289ax2.set_title('Non-Isothermal PFR with Cooling', fontsize=12)290ax2.tick_params(axis='y', labelcolor='b')291ax2_temp.tick_params(axis='y', labelcolor='r')292ax2.grid(True, alpha=0.3)293294# Combine legends295lines1, labels1 = ax2.get_legend_handles_labels()296lines2, labels2 = ax2_temp.get_legend_handles_labels()297ax2.legend(lines1 + lines2, labels1 + labels2, fontsize=10, loc='center right')298299plt.tight_layout()300plt.savefig('reaction_engineering_plot3.pdf', dpi=150, bbox_inches='tight')301plt.close()302303# Calculate volumes for 90% conversion304V_pfr_90 = (F_A0 / k_pfr) * np.log(1 / 0.1)305V_cstr_90 = (0.9 / (k_pfr * 0.1)) * Q_pfr306volume_ratio = V_cstr_90 / V_pfr_90307T_max = np.max(T_profile)308T_rise = T_max - T_inlet309\end{pycode}310311\begin{figure}[H]312\centering313\includegraphics[width=0.95\textwidth]{reaction_engineering_plot3.pdf}314\caption{PFR analysis showing superior performance over CSTR for positive-order reactions. For 90\% conversion, PFR requires $V = \py{f'{V_pfr_90:.1f}'}$ L while CSTR needs $\py{f'{V_cstr_90:.1f}'}$ L, a $\py{f'{volume_ratio:.1f}'}$× difference. The non-isothermal case demonstrates temperature rise of $\py{f'{T_rise:.1f}'}$ K due to exothermic reaction ($\Delta H_R = -50$ kJ/mol), requiring active cooling to prevent thermal runaway.}315\end{figure}316317\section{Arrhenius Temperature Dependence}318319The Arrhenius equation describes the temperature dependence of the rate constant:320\begin{equation}321k(T) = A \exp\left(-\frac{E_a}{RT}\right)322\end{equation}323324where $A$ is the pre-exponential factor, $E_a$ is the activation energy, $R$ is the gas constant, and $T$ is absolute temperature.325326Taking the logarithm yields a linear form:327\begin{equation}328\ln k = \ln A - \frac{E_a}{R} \cdot \frac{1}{T}329\end{equation}330331This enables determination of $E_a$ from the slope of an Arrhenius plot ($\ln k$ vs. $1/T$).332333\begin{pycode}334# Arrhenius temperature dependence analysis335A_arr = 1e10 # pre-exponential factor (1/min)336Ea_arr = 75000 # activation energy (J/mol)337R_const = 8.314 # J/(mol·K)338339# Temperature range340T_range_C = np.linspace(0, 100, 200)341T_range_K = T_range_C + 273.15342343# Calculate rate constants344k_values = A_arr * np.exp(-Ea_arr / (R_const * T_range_K))345346# Reference temperatures347T_ref = 298.15 # 25°C348T_high = 373.15 # 100°C349k_ref = A_arr * np.exp(-Ea_arr / (R_const * T_ref))350k_high = A_arr * np.exp(-Ea_arr / (R_const * T_high))351k_ratio = k_high / k_ref352353# Arrhenius plot data for parameter estimation354T_expt = np.array([273, 298, 323, 348, 373]) # K355k_expt = A_arr * np.exp(-Ea_arr / (R_const * T_expt))356# Add some noise to simulate experimental data357np.random.seed(42)358k_expt_noisy = k_expt * (1 + 0.05 * np.random.randn(len(k_expt)))359360# Linear regression for Ea determination361inv_T = 1 / T_expt362ln_k = np.log(k_expt_noisy)363coeffs = np.polyfit(inv_T, ln_k, 1)364Ea_fitted = -coeffs[0] * R_const365A_fitted = np.exp(coeffs[1])366ln_k_fit = np.polyval(coeffs, inv_T)367368fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))369370# Rate constant vs temperature371ax1.semilogy(T_range_C, k_values, 'b-', linewidth=2)372ax1.plot([T_ref-273.15, T_high-273.15], [k_ref, k_high], 'ro', markersize=8, label='Reference points')373ax1.set_xlabel('Temperature (°C)', fontsize=11)374ax1.set_ylabel('Rate Constant $k$ (min$^{-1}$)', fontsize=11)375ax1.set_title('Arrhenius Temperature Dependence', fontsize=12)376ax1.legend(fontsize=10)377ax1.grid(True, alpha=0.3, which='both')378379# Arrhenius plot for parameter estimation380ax2.plot(inv_T * 1000, ln_k, 'o', markersize=8, label='Experimental data')381ax2.plot(inv_T * 1000, ln_k_fit, 'r--', linewidth=2, label=f'Fit: $E_a$ = {Ea_fitted/1000:.1f} kJ/mol')382ax2.set_xlabel('1000/T (K$^{-1}$)', fontsize=11)383ax2.set_ylabel('$\\ln(k)$', fontsize=11)384ax2.set_title('Arrhenius Plot for Parameter Estimation', fontsize=12)385ax2.legend(fontsize=10)386ax2.grid(True, alpha=0.3)387388plt.tight_layout()389plt.savefig('reaction_engineering_plot4.pdf', dpi=150, bbox_inches='tight')390plt.close()391392# Calculate reaction time sensitivity393X_target_arr = 0.90394t_25C = -np.log(1 - X_target_arr) / k_ref395t_50C = -np.log(1 - X_target_arr) / (A_arr * np.exp(-Ea_arr / (R_const * 323.15)))396time_reduction = (t_25C - t_50C) / t_25C * 100397\end{pycode}398399\begin{figure}[H]400\centering401\includegraphics[width=0.95\textwidth]{reaction_engineering_plot4.pdf}402\caption{Arrhenius analysis showing exponential temperature dependence of reaction rate. With $E_a = \py{f'{Ea_arr/1000:.0f}'}$ kJ/mol, the rate constant increases $\py{f'{k_ratio:.1f}'}$× from 25°C to 100°C. For batch operation achieving 90\% conversion, increasing temperature from 25°C to 50°C reduces reaction time by $\py{f'{time_reduction:.1f}'}$\%, demonstrating the strong coupling between kinetics and thermal management in reactor design.}403\end{figure}404405\section{Levenspiel Plots for Reactor Comparison}406407The Levenspiel plot displays $1/(-r_A)$ versus conversion $X_A$, enabling graphical reactor design and comparison \cite{Levenspiel1999}. The area under the curve represents the required reactor volume:408409For PFR: $V = F_{A0} \int_0^{X_A} \frac{dX_A}{-r_A}$ (area under curve)410411For CSTR: $V = F_{A0} \frac{X_A}{-r_A}$ (area of rectangle)412413This graphical method reveals that PFRs always require less volume than CSTRs for positive-order reactions.414415\begin{pycode}416# Levenspiel plots for different reaction orders417F_A0_lev = 10.0 # molar flow rate (mol/min)418C_A0_lev = 1.0 # inlet concentration (mol/L)419X_vals = np.linspace(0.01, 0.95, 200)420421# First-order reaction422k1_lev = 0.1 # 1/min423C_A_first = C_A0_lev * (1 - X_vals)424r_A_first = -k1_lev * C_A_first425inv_r_A_first = -1 / r_A_first426427# Second-order reaction428k2_lev = 0.15 # L/(mol·min)429C_A_second = C_A0_lev * (1 - X_vals)430r_A_second = -k2_lev * C_A_second**2431inv_r_A_second = -1 / r_A_second432433# Zero-order reaction434k0_lev = 0.1 # mol/(L·min)435r_A_zero = -k0_lev436inv_r_A_zero = -1 / r_A_zero * np.ones_like(X_vals)437438# Calculate volumes for X = 0.80439X_design = 0.80440idx_design = np.argmin(np.abs(X_vals - X_design))441442# PFR volumes (integration)443V_pfr_first = F_A0_lev * np.trapz(inv_r_A_first[:idx_design], X_vals[:idx_design])444V_pfr_second = F_A0_lev * np.trapz(inv_r_A_second[:idx_design], X_vals[:idx_design])445446# CSTR volumes (rectangle)447V_cstr_first = F_A0_lev * X_design * inv_r_A_first[idx_design]448V_cstr_second = F_A0_lev * X_design * inv_r_A_second[idx_design]449450fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))451452# Levenspiel plot453ax1.plot(X_vals, inv_r_A_zero, 'g-', linewidth=2, label='Zero-order')454ax1.plot(X_vals, inv_r_A_first, 'b-', linewidth=2, label='First-order')455ax1.plot(X_vals, inv_r_A_second, 'r-', linewidth=2, label='Second-order')456ax1.axvline(x=X_design, color='gray', linestyle=':', alpha=0.5)457ax1.set_xlabel('Conversion $X_A$', fontsize=11)458ax1.set_ylabel('$-1/r_A$ (min·L/mol)', fontsize=11)459ax1.set_title('Levenspiel Plot for Reactor Comparison', fontsize=12)460ax1.legend(fontsize=10)461ax1.grid(True, alpha=0.3)462ax1.set_ylim(0, 80)463464# Volume comparison465reactor_types = ['PFR\n(1st)', 'CSTR\n(1st)', 'PFR\n(2nd)', 'CSTR\n(2nd)']466volumes = [V_pfr_first, V_cstr_first, V_pfr_second, V_cstr_second]467colors = ['blue', 'lightblue', 'red', 'lightcoral']468469bars = ax2.bar(reactor_types, volumes, color=colors, alpha=0.7, edgecolor='black')470ax2.set_ylabel('Required Volume (L)', fontsize=11)471ax2.set_title(f'Reactor Volume for $X_A = {X_design}$', fontsize=12)472ax2.grid(True, alpha=0.3, axis='y')473474# Add value labels on bars475for bar, vol in zip(bars, volumes):476height = bar.get_height()477ax2.text(bar.get_x() + bar.get_width()/2., height,478f'{vol:.1f}', ha='center', va='bottom', fontsize=10)479480plt.tight_layout()481plt.savefig('reaction_engineering_plot5.pdf', dpi=150, bbox_inches='tight')482plt.close()483484# Calculate efficiency ratios485pfr_cstr_ratio_first = V_cstr_first / V_pfr_first486pfr_cstr_ratio_second = V_cstr_second / V_pfr_second487\end{pycode}488489\begin{figure}[H]490\centering491\includegraphics[width=0.95\textwidth]{reaction_engineering_plot5.pdf}492\caption{Levenspiel plot demonstrating graphical reactor design method. The area under each curve represents PFR volume, while CSTR volume corresponds to the rectangular area. For first-order kinetics at $X_A = \py{X_design}$, CSTR requires $\py{f'{pfr_cstr_ratio_first:.2f}'}$× more volume than PFR. Second-order reactions show even greater disparity ($\py{f'{pfr_cstr_ratio_second:.2f}'}$×), while zero-order reactions perform identically in both configurations.}493\end{figure}494495\section{Reactor Performance Optimization}496497The optimal reactor configuration depends on multiple factors:498\begin{itemize}499\item \textbf{Reaction order}: PFR advantage increases with positive reaction order500\item \textbf{Conversion target}: Higher conversions favor PFR over CSTR501\item \textbf{Heat management}: Exothermic reactions may require staged cooling502\item \textbf{Mixing requirements}: Fast reactions benefit from CSTR back-mixing503\item \textbf{Capital vs operating costs}: Multiple small CSTRs vs single large PFR504\end{itemize}505506\begin{pycode}507# Comprehensive reactor performance comparison508conversion_targets = np.array([0.50, 0.70, 0.80, 0.90, 0.95])509k_opt = 0.1 # rate constant (1/min)510Q_opt = 10.0 # flow rate (L/min)511C_A0_opt = 1.0 # inlet concentration (mol/L)512F_A0_opt = Q_opt * C_A0_opt513514# Calculate required volumes for each configuration515V_pfr_opt = []516V_cstr_1_opt = []517V_cstr_3_opt = []518519for X in conversion_targets:520# PFR521V_pfr = (F_A0_opt / k_opt) * np.log(1 / (1 - X))522V_pfr_opt.append(V_pfr)523524# Single CSTR525V_cstr = (F_A0_opt * X) / (k_opt * (1 - X))526V_cstr_1_opt.append(V_cstr)527528# 3 CSTRs in series529def eq_3cstr(tau):530return 1 - 1/(1 + k_opt*tau/3)**3 - X531tau_3 = brentq(eq_3cstr, 0, 100)532V_cstr_3_opt.append(tau_3 * Q_opt)533534V_pfr_opt = np.array(V_pfr_opt)535V_cstr_1_opt = np.array(V_cstr_1_opt)536V_cstr_3_opt = np.array(V_cstr_3_opt)537538fig, (ax1, ax2) = plt.subplots(1, 2, figsize=(12, 5))539540# Volume comparison541ax1.plot(conversion_targets * 100, V_pfr_opt, 'b-o', linewidth=2, markersize=6, label='PFR')542ax1.plot(conversion_targets * 100, V_cstr_1_opt, 'r--s', linewidth=2, markersize=6, label='Single CSTR')543ax1.plot(conversion_targets * 100, V_cstr_3_opt, 'g-.^', linewidth=2, markersize=6, label='3 CSTRs in series')544ax1.set_xlabel('Conversion (\\%)', fontsize=11)545ax1.set_ylabel('Required Volume (L)', fontsize=11)546ax1.set_title('Reactor Volume vs Conversion', fontsize=12)547ax1.legend(fontsize=10)548ax1.grid(True, alpha=0.3)549550# Efficiency ratios (CSTR/PFR)551ratio_cstr_1 = V_cstr_1_opt / V_pfr_opt552ratio_cstr_3 = V_cstr_3_opt / V_pfr_opt553554ax2.plot(conversion_targets * 100, ratio_cstr_1, 'r--s', linewidth=2, markersize=6, label='Single CSTR / PFR')555ax2.plot(conversion_targets * 100, ratio_cstr_3, 'g-.^', linewidth=2, markersize=6, label='3 CSTRs / PFR')556ax2.axhline(y=1.0, color='b', linestyle='-', alpha=0.5, label='PFR baseline')557ax2.set_xlabel('Conversion (\\%)', fontsize=11)558ax2.set_ylabel('Volume Ratio (relative to PFR)', fontsize=11)559ax2.set_title('Reactor Efficiency Comparison', fontsize=12)560ax2.legend(fontsize=10)561ax2.grid(True, alpha=0.3)562563plt.tight_layout()564plt.savefig('reaction_engineering_plot6.pdf', dpi=150, bbox_inches='tight')565plt.close()566567# Store key optimization metrics568V_90_pfr = V_pfr_opt[-2]569V_90_cstr1 = V_cstr_1_opt[-2]570V_90_cstr3 = V_cstr_3_opt[-2]571savings_3cstr = (V_90_cstr1 - V_90_cstr3) / V_90_cstr1 * 100572\end{pycode}573574\begin{figure}[H]575\centering576\includegraphics[width=0.95\textwidth]{reaction_engineering_plot6.pdf}577\caption{Reactor optimization analysis across conversion targets. The efficiency gap between CSTR and PFR widens dramatically at high conversions - at 90\% conversion, single CSTR requires $\py{f'{V_90_cstr1:.1f}'}$ L versus $\py{f'{V_90_pfr:.1f}'}$ L for PFR. Staging with 3 CSTRs in series reduces volume to $\py{f'{V_90_cstr3:.1f}'}$ L, achieving $\py{f'{savings_3cstr:.1f}'}$\% savings while preserving mixing benefits for heat management and fast reactions.}578\end{figure}579580\section{Results Summary}581582\begin{pycode}583# Compile key results from all analyses584results_data = [585['Batch (1st-order)', f'{t_90_first:.1f}', 'min', 'Time for 90\\% conversion'],586['Batch (2nd-order)', f'{t_90_second:.1f}', 'min', 'Time for 90\\% conversion'],587['CSTR (single)', f'{tau_single:.1f}', 'min', 'Residence time for 90\\%'],588['CSTR staging reduction', f'{reduction_pct:.1f}', '\\%', '3 stages vs single'],589['PFR volume (90\\%)', f'{V_pfr_90:.1f}', 'L', 'First-order reaction'],590['CSTR/PFR ratio', f'{volume_ratio:.1f}', '--', 'At 90\\% conversion'],591['Arrhenius $E_a$', f'{Ea_arr/1000:.0f}', 'kJ/mol', 'Activation energy'],592['$k$ ratio (25-100°C)', f'{k_ratio:.1f}', '--', 'Rate constant increase'],593['Time savings (25-50°C)', f'{time_reduction:.1f}', '\\%', 'Batch operation'],594]595596print(r'\begin{table}[H]')597print(r'\centering')598print(r'\caption{Summary of Key Reactor Design Results}')599print(r'\begin{tabular}{@{}llll@{}}')600print(r'\toprule')601print(r'Parameter & Value & Units & Description \\')602print(r'\midrule')603for row in results_data:604print(f"{row[0]} & {row[1]} & {row[2]} & {row[3]} \\\\")605print(r'\bottomrule')606print(r'\end{tabular}')607print(r'\end{table}')608\end{pycode}609610\section{Conclusions}611612This comprehensive computational analysis of chemical reactor design demonstrates several fundamental principles:613614\textbf{Reactor Configuration}: For first-order irreversible reactions at 90\% conversion, PFR achieves residence time of $\tau = \py{f'{V_pfr_90/Q_pfr:.1f}'}$ min, while single CSTR requires $\tau = \py{f'{tau_single:.1f}'}$ min - a $\py{f'{volume_ratio:.1f}'}$× difference. CSTR staging (3 reactors in series) reduces this gap by $\py{f'{reduction_pct:.1f}'}$\%, approaching PFR efficiency while maintaining mixing advantages.615616\textbf{Reaction Order Effects}: Levenspiel analysis reveals that higher-order reactions show greater disparity between reactor types. Second-order reactions require $\py{f'{pfr_cstr_ratio_second:.2f}'}$× more CSTR volume than PFR at $X_A = \py{X_design}$, while zero-order kinetics show no preference.617618\textbf{Temperature Management}: Arrhenius behavior ($E_a = \py{f'{Ea_arr/1000:.0f}'}$ kJ/mol) demonstrates that increasing temperature from 25°C to 50°C reduces batch time by $\py{f'{time_reduction:.1f}'}$\%. Non-isothermal PFR analysis shows temperature rise of $\py{f'{T_rise:.1f}'}$ K for exothermic reactions, requiring active cooling to prevent thermal runaway and maintain selectivity.619620\textbf{Design Optimization}: The optimal configuration balances capital costs (reactor volume), operating costs (temperature control), and process requirements (conversion, selectivity). Staged CSTR systems offer a practical compromise, achieving $\py{f'{savings_3cstr:.1f}'}$\% volume reduction versus single CSTR while preserving thermal management flexibility for fast exothermic reactions.621622These numerical methods provide essential tools for reactor scale-up, process intensification, and economic optimization in chemical manufacturing \cite{Fogler2016, Schmidt2005}.623624\begin{thebibliography}{99}625626\bibitem{Fogler2016}627Fogler, H. S. (2016). \textit{Elements of Chemical Reaction Engineering} (5th ed.). Prentice Hall.628629\bibitem{Levenspiel1999}630Levenspiel, O. (1999). \textit{Chemical Reaction Engineering} (3rd ed.). John Wiley \& Sons.631632\bibitem{Smith2005}633Smith, J. M., Van Ness, H. C., \& Abbott, M. M. (2005). \textit{Introduction to Chemical Engineering Thermodynamics} (7th ed.). McGraw-Hill.634635\bibitem{Scott2006}636Scott, D. M. (2006). \textit{Industrial Process Systems}. Springer.637638\bibitem{Schmidt2005}639Schmidt, L. D. (2005). \textit{The Engineering of Chemical Reactions} (2nd ed.). Oxford University Press.640641\bibitem{Aris1989}642Aris, R. (1989). \textit{Elementary Chemical Reactor Analysis}. Butterworth-Heinemann.643644\bibitem{Davis2003}645Davis, M. E., \& Davis, R. J. (2003). \textit{Fundamentals of Chemical Reaction Engineering}. McGraw-Hill.646647\bibitem{Froment2011}648Froment, G. F., Bischoff, K. B., \& De Wilde, J. (2011). \textit{Chemical Reactor Analysis and Design} (3rd ed.). John Wiley \& Sons.649650\bibitem{Carberry2001}651Carberry, J. J. (2001). \textit{Chemical and Catalytic Reaction Engineering}. Dover Publications.652653\bibitem{Hill1977}654Hill, C. G. (1977). \textit{An Introduction to Chemical Engineering Kinetics and Reactor Design}. John Wiley \& Sons.655656\bibitem{Walas1991}657Walas, S. M. (1991). \textit{Modeling with Differential Equations in Chemical Engineering}. Butterworth-Heinemann.658659\bibitem{Denbigh1981}660Denbigh, K. G., \& Turner, J. C. R. (1981). \textit{Chemical Reactor Theory: An Introduction} (3rd ed.). Cambridge University Press.661662\bibitem{Holland1975}663Holland, C. D., \& Anthony, R. G. (1975). \textit{Fundamentals of Chemical Reaction Engineering}. Prentice Hall.664665\bibitem{Westerterp1984}666Westerterp, K. R., Van Swaaij, W. P. M., \& Beenackers, A. A. C. M. (1984). \textit{Chemical Reactor Design and Operation} (2nd ed.). John Wiley \& Sons.667668\bibitem{Nauman2008}669Nauman, E. B. (2008). \textit{Chemical Reactor Design, Optimization, and Scaleup} (2nd ed.). John Wiley \& Sons.670671\bibitem{Rawlings2002}672Rawlings, J. B., \& Ekerdt, J. G. (2002). \textit{Chemical Reactor Analysis and Design Fundamentals}. Nob Hill Publishing.673674\bibitem{Missen1999}675Missen, R. W., Mims, C. A., \& Saville, B. A. (1999). \textit{Introduction to Chemical Reaction Engineering and Kinetics}. John Wiley \& Sons.676677\bibitem{Butt2000}678Butt, J. B. (2000). \textit{Reaction Kinetics and Reactor Design} (2nd ed.). Marcel Dekker.679680\bibitem{Trambouze2004}681Trambouze, P., Van Landeghem, H., \& Wauquier, J. P. (2004). \textit{Chemical Reactors: Design, Engineering, Operation}. Editions Technip.682683\bibitem{Ancheyta2017}684Ancheyta, J. (2017). \textit{Chemical Reaction Kinetics: Concepts, Methods and Case Studies}. John Wiley \& Sons.685686\end{thebibliography}687688\end{document}689690691