Open Access Article
Eleni Grammenou
a,
Andrea Friso
a,
Simon Dawes
b,
Nicholas Snead
a,
Georgios Gkogkos
a,
Maximilian O. Besenhard
a,
Maria Mourkou
a,
Federico Galvanin
*a and
Asterios Gavriilidis
*a
aDepartment of Chemical Engineering, University College London, Torrington Place, London, WC1E 7JE, UK. E-mail: f.galvanin@ucl.ac.uk; a.gavriilidis@ucl.ac.uk
bWorkshop of Biochemical Engineering, Chemical Engineering and Physics, University College London, Torrington Place, London, WC1E 7JE, UK
First published on 8th September 2025
Magnetically agitated miniaturised continuous stirred tank reactors (m-CSTRs) are an attractive tool for the investigation of reaction kinetics, as they combine active stirring with enhanced heat and mass transfer due to their small dimensions, while their compatibility with in situ spectroscopic techniques enables reaction monitoring and high-throughput data acquisition. This study presents the development of a 2.65 mL m-CSTR, integrated with in situ Raman spectroscopy for real-time kinetic data acquisition in continuous flow. The reactor, featuring a temperature-controlled stainless-steel chamber with a top quartz glass window and a PTFE slotted impeller, was assessed for its macro- and micromixing characteristics at flowrates between 0.5 and 4 mL min−1. The slotted impeller led to near-ideal CSTR behaviour and improved micromixing quality in comparison to conventional cross stir bars of similar dimensions. Using the imine synthesis of n-benzylidenebenzylamine from benzaldehyde and benzylamine as a model reaction, kinetic parameters were determined with direct composition measurement inside the reactor and the most informative region of the design space, i.e., the experimental conditions that provide the most useful data for accurate kinetic parameter estimation and model validation, was identified to be in the region of high reactant concentrations and short residence times within the temperature range investigated (15–45 °C).
Raman spectroscopy allows fast, contact-free, non-destructive sampling with high chemical specificity. Importantly, water does not produce Raman signals in the spectral regions where organic compounds typically exhibit peaks, enabling organics detection without water interference.14 Although traditionally associated with qualitative analysis, Raman spectroscopy is effective in quantitative applications by leveraging the proportionality between the intensity of an analyte band and its concentration.15 The recent advances in Raman process spectrometers, equipped with flow cells and contactless probes, enhance their seamless integration into flow processes. Schwolow et al.16 demonstrated the use of a contactless Raman probe to study the kinetics of 3-piperidino-propionic acid ethyl ester synthesis in an aluminium plate microreactor under steady and transient conditions. Cao et al.17 employed a silicon-glass microreactor system with four micro-packed beds to investigate the multiphase oxidation of benzyl alcohol, using a portable Raman spectrometer, thereby highlighting the application of operando Raman spectroscopy for catalyst evaluation. Both studies demonstrate the use of intricate channel geometries in microfluidic chip reactors, which allow for measurement at multiple locations, a feature particularly useful for fast reactions. However, for reactions requiring longer residence times, necessary adjustments to the microchannel lengths could result in increased pressure drop and extended time and material consumption to reach steady state. To address this challenge, Klement et al.18 introduced a novel sandwich minichannel reactor for oscillating droplets that enabled Raman spectroscopy to analyse individual droplets within a PTFE channel. The same group further streamlined this reactor design, applying it to study the kinetics of acrylic acid polymerisation in slug flow.19 Schulz et al.20 also employed this reactor setup to investigate the kinetics of imine synthesis, coupling Raman spectroscopy with chemometrics.
Miniaturised continuous stirred tank reactors (m-CSTRs) provide an alternative that merges the small footprint and reagent efficiency of microreactors with the active mixing and back-mixing characteristics of conventional CSTRs. With volumes of a few millilitres, m-CSTRs are well-suited for solid-producing reactions,21–23 synthesis of nanoparticles,24–26 multiphasic flow chemistry27,28 and investigation of slower reactions.29 A key advantage of m-CSTRs lies in their operational versatility, an aspect that is particularly useful for reaction kinetic studies. When operated in continuous flow, they enable the collection of steady-state data under precisely controlled conditions. Incorporating transparent windows allows for integration with non-invasive spectroscopic tools, such as in situ Raman (using quartz windows) and transflectance IR spectroscopy (using CaF2 windows). This capability would be valuable for high-throughput kinetic screening, automated experimentation, and process development.30–32 In contrast to many microreactor formats that are limited to continuous flow, m-CSTRs can also be adapted for batch operation with minimal modifications. However, batch operation in conjunction with real-time spectroscopic monitoring would generally be better suited for slower reaction systems. For fast reactions, a significant portion of conversion can occur during the initial filling or mixing phase before accurate sampling and data collection can begin. This “dead time” at the start of the experiment would make it difficult to accurately capture the initial kinetics with spectroscopic techniques. This issue may not arise in all systems; Guan et al.33 reported batch studies on multiphasic reactions with online pressure measurement. However, offline analysis was ultimately preferred for accurate concentration measurements.
Despite the clear promise of m-CSTRs for kinetic investigations, especially in continuous operation, no prior reports have demonstrated their use for acquiring continuous flow kinetic data using in situ spectroscopy. Our work introduces a temperature-controlled m-CSTR, equipped with a Raman spectrometer probe that analyses the reaction mixture inside the reactor, for real-time kinetic data acquisition. The combination of in situ spectroscopic monitoring and improved mixing performance at this scale, through a custom-made slotted impeller, allows our reactor to collect accurate kinetic data under steady-state operation, demonstrating a pathway for automating and streamlining kinetic analysis in flow chemistry. We begin by evaluating the mixing behaviour of the m-CSTR, assessing both micromixing and macromixing quality. We then demonstrate the reactor's capability to rapidly generate kinetic data through integration with an automated experimental platform. To validate our system, we selected a fast, homogeneous liquid-phase imine synthesis as a model reaction. This choice presents stringent demands on mixing efficiency and time-resolved spectral acquisition, providing a rigorous proof of concept for the reactor performance. Successful kinetic monitoring of this rapid reaction demonstrates the robustness of our approach and establishes a foundation for future studies involving more complex or multiphase systems.
1. Reactor chamber: Constructed from 316 stainless steel, the cylindrical chamber has an internal diameter of 15 mm and a depth of 15 mm, resulting in a total internal volume of 2.65 mL.
2. Optical window: A 25 mm diameter, 2 mm thick quartz optical grade window (Knight Optical) is mounted on top of the reactor chamber, providing optical access for the Raman probe.
3. Reactor lid: A 316 stainless steel lid seals the reactor chamber together with the optical window, ensuring a secure, leak-proof environment.
4. Cooling/heating jacket: Made of 316 stainless steel, the jacket has a square body (90 × 90 × 24.65 mm) with an inner protruding fin to enhance the recirculation of cooling/heating fluid.
5. Reactor cover: A 316 stainless steel reactor cover is used to minimise heat loss due to convection with the surrounding environment, maintaining stable temperature conditions within the reactor.
Two 2 mm I.D. holes were drilled at a 90° angle to each other on the side walls in close proximity to the bottom of the reactor chamber, designed for the introduction of the reactants. A 2 mm I.D. hole, used as the outlet, was drilled on the side wall, 4 mm below the top of the reactor chamber, to minimise short-circuiting of the reacting mixture. A custom-made slotted impeller with dimensions of 8 mm in height and 13 mm in diameter was designed and fabricated from PTFE. The impeller featured a pocket for fitting a high-pull rare earth magnet to enable coupling with an external magnetic stirring unit to control the impeller's rotational speed. Its asymmetric design consisted of four arms containing a set of fins, separated by narrow slots (gaps) to enhance mixing. The volume of the impeller without the magnet was 0.24 mL and 0.30 mL including the magnet. As such, when inserted into the reactor (total internal volume: 2.65 mL), the effective working volume was reduced to 2.35 mL. All the reactor components, including the impeller, were machined using CNC milling (Haas TM1P Vertical Machining Centre). The whole unit has the capability to withstand harsh chemical environments, and the integration of a heating jacket allows conducting reactions at the desired temperature by connecting it with a circulating bath. Further details on the design and fabrication of the reactor system can be found in the SI, S1. Fig. 1 shows the CAD model, a picture of the fully assembled reactor and an exploded view of the reactor. For the investigation of micromixing performance of the PTFE custom-made slotted impeller, a PTFE cross-stir bar of 10 mm in diameter and 5 mm in height (total volume of 0.18 mL) was also employed for comparison.
![]() | ||
| Fig. 1 (a) CAD design of m-CSTR, (b) exploded reactor assembly, with a zoomed-in detail of the custom-designed slotted impeller and (c) picture of assembled m-CSTR. | ||
The concentration was normalised, based on the maximum concentration of the tracer injected. Short sections of tubing were used to connect the inlet of the reactor to the four-way valve and the outlet of the reactor to the flow cell. Despite their relatively short lengths, the outlet concentration profile of the system was a convolution of the RTD of the reactor and the connecting tubings. To obtain only the m-CSTR's RTD, an additional RTD experiment was performed connecting the inlet and outlet tubing sections directly, to evaluate their contribution to the system's output. Fast Fourier transform (FFT) and smoothening of the data using the Savitzky–Golay filter were used for deconvolution of the signal.35 A detailed description of the data post-treatment can be found in Gkogkos et al.25 and the SI (section S5). The normalised deconvoluted RTD function of the reactor Eθ was compared with that of the ideal CSTR model Eθ,model (eqn (1)) by evaluating the coefficient of determination, R2, between the experimental and the model RTD curves (eqn (2)).
| Eθ,model = e−θ | (1) |
![]() | (2) |
θ is the average of the experimental values Eθ over the range θ = 0 − 4.5.
| H2BO3− + H+ ↔ H3BO3 | (3) |
| 5I− + IO3− + 6H+ ↔ 3I2 + 3H2O | (4) |
The acid–base neutralisation reaction (eqn (3)) and the redox reaction (eqn (4)) are competitive reactions for the available protons supplied by the sulfuric acid, with the former being quasi-instantaneous and the latter being rapid, but with a significantly lower reaction rate than that of the neutralisation reaction. When micromixing is not instantaneous, the redox reaction proceeds due to incomplete consumption of protons, resulting in the generation of iodine. The generated iodine from the redox reaction reacts further with excess iodide to reversibly generate triiodide ions, as shown in eqn (5).
| I2 + I− ↔ I3− | (5) |
The triiodide formed has a distinct yellow colour and a strong UV-vis absorption peak at the wavelength of 353 nm and can be quantified using the Beer–Lambert law, as presented in eqn (6).
![]() | (6) |
209 L mol−1 cm−1
37) and loptical the optical path length within the measurement cell.
The degree of micromixing efficiency can be quantitatively described by the segregation index XS, and the mixing time, tm. Both parameters' estimation is relying on the selection of a mixing model, able to describe the mechanistic properties of the integration of the two fluid streams. The segregation index XS is defined as
![]() | (7) |
In this work, the concentrations proposed by Commenge and Falk38 for set 1c were used. The correlation between the mixing time and the triiodide concentration (or equivalently the segregation index, XS) was established by specifying the initial reagent concentrations and solving the Interaction by Exchange with the Mean (IEM)39,40 model for various mixing times (see SI, S6).
For the micromixing analysis, the prepared acid and buffer solutions were loaded separately in 100 mL syringes (SGE) and were pumped simultaneously by the same syringe pump (PHD ULTRA, Harvard) (Fig. 2b). The reactor was stirred by adjusting the rotational speed in a digital hotplate-stirrer (RCT 5 digital, IKA). Flowrates similar to the RTD and reaction kinetic experiments were used, ranging from 0.5 to 4 mL min−1 and stirring speeds of 250, 500, 750 and 1000 rpm were investigated. The outlet of the m-CSTR was connected to the UV-vis analysis set-up by 1 m long PFA tubing (delay loop) to ensure that the reactions were completed and equilibrated. The triiodide concentration was monitored at 353 nm, using the same UV-vis system, as in the RTD experiments (section 2.2.1) but with optical fibres (200–1100 nm, Ocean Optics) optimised for transmission in the UV region. The experiments were repeated three times with the two reactant streams swapped between the inlets, and no differences were observed in the obtained data.
The synthesis of n-benzylidenebenzylamine follows second order kinetics20,41,42 (eqn (8)).
| −rBAl = kCBAlCBAm | (8) |
:
1, allows for the simplification of the reaction rate to eqn (9).| −rBAl = kCBAl2 | (9) |
Fig. 4 provides a simplified schematic representation of the experimental set-up. Two glass syringes (25 mL each, CETONI) were separately filled with a solution of benzaldehyde in methanol and benzylamine in methanol (2 M each). Another glass syringe (25 mL, CETONI) was filled with methanol and acted as the dilution stream for one of the reactants. Dilution of only one of the streams was sufficient to achieve equal inlet concentrations of both reactants. All syringes were mounted on a Nemesys medium-pressure syringe pump (CETONI). The inlet concentrations of benzaldehyde and benzylamine entering the reactor were determined by the relative pump flowrates. Mixing in the reactor was achieved by adjusting the stirring speed of the magnetic hotplate-stirrer. To ensure isothermal operation, the feed streams were preheated using a digital magnetic hotplate-stirrer (RCT 5 digital, IKA), while a cooling/heating circulator (CC-K6, Huber) was connected to the reactor's jacket ports, allowing water recirculation with adjustable temperature. The reaction temperature was controlled by a type K thermocouple (RS PRO) that was inserted into the reaction chamber through a T-junction (IDEX), located at the outlet of the reactor. This thermocouple was connected to an Arduino control box and provided the set-point that was used to adjust the water bath temperature through a PID control script (see SI, S3). The reaction mixture exiting the reactor was collected into a waste vessel (Festo). To prevent evaporation of the solvent, the entire system was pressurised to 1 barg with nitrogen (BOC) using a mass flow controller (SLA5850, Brooks), a gas pressure sensor (40PC, 250 psig, Honeywell) and a back-pressure regulator (K series, 250 psig, Swagelok) connected to the waste vessel. The back pressure regulator ensured that the pressure in the waste vessel was no lower than a certain value, while also working as pressure relief valve, expelling excess gas. One pressure relief valve (U-456, 100 psi, IDEX) was installed right after the reactor outlet to avoid overpressure above 7 bar, as the glass syringes used could withstand up to 10 bar.
The experimental process was automated using the Laboratory Virtual Instrument Engineering Workbench (LabVIEW) environment.43 The LabVIEW code allowed the user to run a set of experiments by setting the temperature, feed concentrations and total flowrate at the user interface panel. For visualisation purposes, the temperature and pressure trends of the system were also displayed in real time in graph charts. The code featured overheating and overpressurisation safety limits, which were set to 80 °C (to prevent evaporation in the water circulator) and 2 barg respectively and would automatically stop the experiment if these values were exceeded.
The concentration of the reactants and products was calculated based on the area of the peak corresponding to the characteristic C
O stretching of benzaldehyde at 1680–1730 cm−1 and C
N stretching of n-benzylidenebenzylamine at 1630–1670 cm−1 (benzylamine does not display distinctive peaks), as shown in Fig. 5. Prior to this, the calibration curves for the respective compounds were obtained (see SI, S4). Since Raman spectra often contain baseline drift and random noise, baseline correction was an important step to extract the correct Raman peak intensities for quantitative analysis. To address this, a MATLAB script containing the airPLS algorithm44 was employed that provided a simple and fast estimation of the baseline, requiring the tuning of only a single parameter to ensure baseline smoothness across all recorded spectra with emphasis on the peaks of interest (see SI, S4). The carbon balance was evaluated in all experiments, yielding an average error of approximately 3% and a maximum error of 5%.
![]() | (10) |
![]() | (11) |
![]() | (12) |
| kref = exp(−KP1) | (13) |
| Eα = KP2 × 104 | (14) |
Here, T is the reaction temperature (K), Tref is the reference temperature (K), calculated as the average value of the upper and lower temperature limits of the experimental conditions and R is the ideal gas constant (8.314 J mol−1 K−1).
The parameters were estimated by maximising the log-likelihood function (see SI, S10). The results of the parameter estimation for KP1 and KP2 were analysed in terms of estimated values and statistics to quantify both the precision of the estimates and model's goodness of fit. Precision was assessed using Student's t-test statistics; the t-value at the 95% confidence level of each kinetic parameter should exceed the tabulated reference t-value given by a Student t-distribution.46 Model fitting quality was evaluated using chi-square (χ2) statistics, where the χ2 value should be lower than the 95% reference value from statistical tables for the correct number of degrees of freedom.47 An information analysis was carried out to quantify the relative amount of information distributed across different experiments, through the evaluation of the Relative Fisher Information (RFI).48,49 This distinction is particularly valuable, as it highlights which data points are the most informative within the experimental framework.
The analysis was performed by first computing a metric derived from the Fisher Information Matrix (FIM), an Nθ × Nθ matrix that quantifies the amount of information that can be obtained from a single experiment. The FIM (Ĥθ) was evaluated using the set of experimental conditions denoted by φ, at the estimated values
of the model parameters:
![]() | (15) |
is the prior Fisher information matrix. ŷi refers to the prediction of the i-th concentration at the sampling time tk.
The RFI was then evaluated based on the FIM.48,49
![]() | (16) |
The standard deviation of the measurement error for each response variable (concentration of benzaldehyde, CBAl and concentration of n-benzylidenebenzylamine, CBIm), was calculated using the method of pooled standard deviation50 from repeated identical experiments to be 0.017 M for benzaldehyde and 0.014 M for n-benzylidenebenzylamine, respectively. These standard deviations were used to calibrate the constant variance model for parameter estimation and model adequacy evaluation (see SI, S10). The measurement error was attributed to the Raman spectrometer used in an open beam configuration that can be considered more sensitive to variations in the surrounding environment than a spectroscopic flow cell.
| Control variable | Lower limit | Upper limit |
|---|---|---|
| Flowrate (mL min−1) | 0.5 | 2 |
| Temperature (°C) | 15 | 45 |
| Inlet concentration of benzaldehyde (M) | 0.5 | 1 |
In all cases, when stirring was applied, the flow field was dominated by the tangential velocity component, which is associated with the rotational motion of the impeller and showed minimal sensitivity to variations in flowrate. Considering that flowrate effects were likely to be most pronounced near the inlet and outlet, the velocity field was closely examined in these regions (see SI, Fig. S11). At a flowrate of 1 mL min−1, the fluid near the inlet was driven back into a localised recirculation zone by the dominant tangential flow. This recirculation zone was absent at the higher flowrate of 4 mL min−1. This effect was consistently observed at stirring speeds of 500 rpm and 1000 rpm, whereas at 250 rpm, the extent of recirculation was significantly diminished. The simulations were repeated with the impeller positioned at two different angles to investigate the discrepancies in the development of the instantaneous velocity distribution with regards to the relative position of the impeller to the inlet (as MRF captures only a moment of the impeller’s rotation), revealing consistent recirculation zones for both angles. The differing behaviour observed with this specific impeller design, compared to the simpler pill shaped magnetic stir bar previously reported, may be explained as follows. It is possible that similar recirculation patterns could occur with stir bars under comparable flowrate conditions. However, in the study of Gkogkos et al.,26 RTD experiments were conducted at flowrates of 2 mL min−1 and then increased directly to 14 mL min−1, a range sufficient to induce some degree of local bypassing. Furthermore, the impeller employed in this study differed significantly from the pill shaped magnetic stir bar. This custom impeller design generated a strong swirling tangential velocity profile near the impeller extremities, which progressively attenuated with distance from the impeller extremities towards the reactor wall, particularly near the top and bottom of the reactor chamber. In contrast, pill shaped stir bars produce a more localised tangential flow. Notably, the slotted impeller spans nearly the entire tank and has a larger diameter (relative to the tank size) compared to the stir bars.
Given the complex hydrodynamics resulting from the confinement of the flow within the reactor, these observations underscore the importance of the impeller design in influencing flow patterns and macromixing efficiency. As the reactor's Eθ curve showed a high degree of agreement with the ideal CSTR model (R2 > 0.90) across the investigated cases, the m-CSTR was modelled as an ideal stirred reactor.
For the micromixing analysis, stirring speed between 250 and 1000 rpm and inlet stream flowrates ranging from 0.25 mL min−1 to 2 mL min−1 (total flowrate Q of 0.5 mL min−1 and 4 mL min−1) were investigated. The slotted impeller and a cross stir bar were tested for water solutions with an acid-to-buffer solution flowrate ratio of 1. In Fig. 7a and b, the micromixing time, tm, calculated by the IEM model39,40 is plotted against the stirring speed for each impeller (the segregation index is reported in the SI, S7). Fig. 7a and b reveal that for both stirring configurations, micromixing time was dependent on the stirring speed, N. As the value of N increased, micromixing time for the slotted impeller and the cross stir bar tended to decrease. This can primarily be attributed to the greater energy input into the fluid and the increased velocity gradient and shearing force within the stirring unit. The decrease was sharper between 250 and 500 rpm and with a less steep gradient for N > 500 rpm for the flowrate range studied, showing less sensitivity to the flowrate at larger N, where the high turbulent kinetic energy dissipation dominated the micromixing process. For both impellers at N = 1000 rpm, an increase in the micromixing time was observed. This increase may result from an antagonistic effect of the inlet flow and the increased stirring rates and can also be associated with the CFD observations in the macromixing analysis. At N = 1000 rpm, the impellers completed 16.7 rotations per second, translating into 1 rotation per 6 milliseconds, a timeframe comparable to the micromixing time measured. At lower flowrates, the entering fluid had reduced kinetic energy. Localised recirculation zones formed near the inlets due to the strong swirling flow generated by the impeller. As a result, reactant molecules did not have sufficient time to diffuse and react. They were rapidly swept away from the reaction zone at the bottom of the reactor, in the space between the inlets and the impeller. This trend was less significant in the case of higher inlet velocities, based on the micromixing values for flowrates of 4 mL min−1, where tm decreased up to 750 rpm and then remained independent of the stirring speed.
For the custom-made slotted impeller, Fig. 7a highlights the effect of the inlet flowrate on the micromixing quality. As the flowrate increased from 0.5 mL min−1 to 4 mL min−1, micromixing times reduced significantly across all stirring speeds, with the highest flowrate of 4 mL min−1 consistently yielding the shortest micromixing times. In contrast, for the cross stir bar (Fig. 7b), the effect of stirring speed dominated over the convection effect of the flowrate, with a sharp increase in the micromixing efficiency evidenced by an average 43% reduction in micromixing time at stirring speeds between 250 and 500 rpm, regardless of the flowrate. Up to 750 rpm, micromixing time remained nearly the same across the range of flowrates studied, with variations within the limits of experimental error, as indicated by the error bars. Overall, the m-CSTR unit employing the slotted impeller demonstrated satisfactory micromixing and displayed better performance across the tested inlet flowrates, compared to the conventional cross stir bar, especially for low stirring speeds.
| Parameter | Value | Standard deviation | 95% t-value |
|---|---|---|---|
| KP1 | 3.47 | 0.05 | 76.63 |
| KP2 | 2.75 | 0.28 | 9.81 |
| Reference t-value (95%) | 1.69 | ||
| χ2 (χref2 = 48.50) | 41.21 |
The adequacy of the 2nd order kinetic model is shown graphically in the parity plot of Fig. 8, where the model predictions for the concentration of benzaldehyde and n-benzylidenebenzylamine fall around the interval of the experimental standard deviation.
![]() | ||
| Fig. 8 Parity plot for model predicted and experimentally measured reactor concentrations of benzaldehyde and n-benzylidenebenzylamine. The dashed lines represent the one-standard deviation limits. | ||
To evaluate the most informative experiments in the design space investigated, the RFI, which represents the fraction of information that each experiment is returning given the parameters KP1 and KP2, was computed for each experimental condition (i.e., 12 different DoE points, the conditions of which can be found in Table S3 of the SI). Based on Fig. 9 that shows the RFI of the average of 3 replicates of 12 steady-state experimental conditions, it is evident that experiment 2 (SS2) yielded the highest RFI value, followed by experiments 6 (SS6) and 10 (SS10). Thus, the experimental region characterised by high reactant concentrations and short residence times (i.e., high flowrate) significantly enhanced the estimation of model parameters. Conversely, experiments 1, 7 and 11 (SS1, SS7, SS11) that were carried out under (mainly) low flowrate and low inlet concentration conditions were less informative. Notably, while SS1 was performed at the lowest temperature (15 °C), SS7 and SS11 correspond to the medium (30 °C) and high (45 °C) range, suggesting that temperature alone is not the determining factor for the low RFI values observed. Rather, it appears that reactant concentration and residence time (flowrate) have a more dominant influence on the information content of the experiments in this system.
![]() | ||
| Fig. 9 RFI evaluation for kinetic experiments in the m-CSTR. For the conditions of each experiment refer to Table S3, SI. | ||
The estimated kinetic parameters of the imine synthesis are compared with reported values in Table 3. Fath et al.42 used a coiled stainless steel microreactor of total inner volume of 1.87 mL with an in situ FTIR spectrometer at the end of the reactor. The kinetic data were estimated using two different approaches: previously determined calibration curves and chemometrics. Based on calibration, a reaction constant of 0.025 L mol−1 s−1 was reported at 25 °C, accompanied by an activation energy of 39.2 kJ mol−1. When applying the Self-Modelling Curve Resolution (SMCR) method, the reaction rate constant varied between 0.026 and 0.065 L mol−1 s−1, while the activation energy ranged from 41.6 kJ mol−1 to 43.3 kJ mol−1. Schulz et al.20 combined a nearly isothermal oscillating segmented flow reactor (where each 7 μL droplet behaved practically as a batch reactor) and Raman spectroscopy. Similarly to Fath et al.,42 chemometrics were employed to obtain calibration-free kinetic parameters. They reported a reaction rate coefficient of 0.036–0.037 L mol−1 s−1 at 25 °C and an activation energy in the range of 27.4–29.3 kJ mol−1. The kinetic rate constant and activation energy obtained in our work are of similar magnitude to the ones reported by Fath et al.42 and Schulz et al.20
| Reactor setup | Mode | Temperature (°C) | Residence time (min) | Inlet benzaldehyde concentration (M) | kref at 25 °C (L mol−1 s−1) | Activation energy (kJ mol−1) |
|---|---|---|---|---|---|---|
| Note: the asterisk (*) indicates reaction time rather than residence time. | ||||||
| Coiled stainless steel microreactor (Fath et al.42) | Steady-state | 15–35 | 0.2–6 | 2 | 0.025–0.065 | 39.2–43.3 |
| Oscillating segmented flow reactor (Schulz et al.20) | Stopped-flow | 15–45 | 0–5* | 0.5–1 | 0.036–0.037 | 27.4–29.3 |
| m-CSTR (this work) | Steady-state | 15–45 | 1.2–4.6 | 0.5–1 | 0.026 | 27.5 |
Data for this article, including micromixing times and Raman-measured concentrations, are available at the Science Data Bank (https://doi.org/10.57760/sciencedb.22997). The code is publicly available at: https://github.com/AndreaFriso/Design-characterisation-and-application-of-a-miniaturised-CSTR.
| This journal is © The Royal Society of Chemistry 2025 |