Abstract
The current throughput of conventional organic chemical synthesis is usually a few experiments for each operator per day. We develop a robotic system for ultra-high-throughput chemical synthesis, online characterization, and large-scale condition screening of photocatalytic reactions, based on the liquid-core waveguide, microfluidic liquid-handling, and artificial intelligence techniques. The system is capable of performing automated reactant mixture preparation, changing, introduction, ultra-fast photocatalytic reactions in seconds, online spectroscopic detection of the reaction product, and screening of different reaction conditions. We apply the system in large-scale screening of 12,000 reaction conditions of a photocatalytic [2 + 2] cycloaddition reaction including multiple continuous and discrete variables, reaching an ultra-high throughput up to 10,000 reaction conditions per day. Based on the data, AI-assisted cross-substrate/photocatalyst prediction is conducted.
Subject terms: Flow chemistry, Automation, Synthetic chemistry methodology
The current throughput of conventional organic chemical synthesis is usually a few experiments for each operator per day. Here the authors develop a robotic system for ultra-high-throughput chemical synthesis, online characterization and large-scale condition screening of photocatalytic reactions.
Introduction
How many chemical synthesis reactions can an organic synthesizer complete in a day? In most organic synthesis laboratories around the world, the numbers are usually no more than 10 reactions per day. Chemical synthesis, especially for many organic synthesis reactions, are often time-consuming and take hours or even days to complete. Therefore, the optimization and screening of organic synthesis often requires a significant investment of time and effort by researchers1. Recently, various automated and high-throughput organic synthesis and screening techniques have been developed based on automated robotics and microfluidic chemistry techniques. Flow microreactors have advantages of fast reaction speed and high reaction efficiency due to the scale effect of high mass and heat transfer efficiency in microfluidic channels2–4. Photocatalytic reactions are a type of reactions frequently-performed in organic synthesis that use light irradiation to excite reactant molecules or catalysts to promote the reaction, converting light energy into chemical energy5,6. Compared with the conventional batch photocatalytic reactors, the flow photocatalytic microreactors could reduce the reaction time from several days or a few hours to several hours or even minutes7,8. However, in spite of the significant improvement in reaction speed of the current flow photocatalytic systems, their throughputs are still much lower than what is needed for the application of the big data-based artificial intelligence (AI) technique to the field of chemical synthesis9–14, which has become one of the current hotspots of chemical research as it is considered to have the potential to transformationally improve the efficiency of chemical synthesis research and development15–18. The lack of a large amount of reliable and high-quality data is also the major factor currently hindering the application of AI technique in chemical synthesis19–25. To address this challenge, we developed an automated high-throughput system that can complete ultra-fast photocatalytic reactions in the seconds scale and ultra-large scale screening up to 10,000 reactions per day, using the microfluidic liquid-core waveguide (LCW)26,27, automated microfluidic liquid-handling and AI techniques.
Results
Ultra-fast photocatalytic synthesis and characterization
We designed and constructed a new type of microfluidic photocatalytic microreactor using the LCW technique to introduce high-intensity laser light into the microfluidic photocatalytic reaction channel to significantly increase the speed of photocatalytic reaction (Fig. 1A). We used four 450-nm lasers as light sources and coupled a total of ca. 4.6 W of light into the reaction area through four optical fibers. The light transmitted by multiple total reflections in the 5-cm-long reaction channel based on the LCW principle, resulting in a laser luminous flux over 3.5 × 104 mW/cm2 through the channel cross section within a channel length range of 3 cm (Fig. 1B). Such a light irradiation intensity is ca. 10,000 times more than those in conventional batch photocatalytic systems28,29. In the microfluidic LCW microreactor, the laser light was reflected multiple times in the photocatalytic microreactor, resulting in a more uniform high-intensity light distribution in the 3-cm-length middle part of the 530-μm-diameter reaction channel, which was 15 times more than the ca. 2 mm irradiation depth in conventional batch photocatalytic systems. However, we observed that the use of the high-intensity laser irradiation caused a significant increase in temperature in the capillary reactor, resulting in an increase in side reactions and a decrease in reaction yield. To address this problem, we designed the microfluidic LCW photocatalytic microreactor as a cannula configuration with an inner reaction capillary and an outer capillary with the circulating cooling solution flowing through to control the temperature of the reaction channel, for instance, 25 ± 2 °C.
We applied the LCW photocatalytic microreactor to a typical organic photocatalytic synthesis, photocatalytic [2 + 2] cycloaddition reaction30, which has gained extensive attention because cyclobutane obtained through this reaction has been frequently found in a variety of natural products with biological activity31–33. Using the LCW photocatalytic microreactor, the photocatalytic [2 + 2] cycloaddition reaction of substrate S-1 could be completely converted with a residence time of only 3.3 s in the microreactor channel, while the reaction yield and diastereomeric ratio (d.r.) were comparable to those of conventional batch photocatalytic systems. As a comparison, the time consumption of this photocatalytic [2 + 2] cycloaddition reaction in a conventional batch photocatalytic reactor was up to 4 h with the same photocatalyst and substrate species, concentration and photocatalyst ratio conditions. Our LCW photocatalytic microreactor could shorten the reaction time by 4300 times with the aid of ultra-high light intensity, uniform and long-length light irradiation, stable reactor temperature control and microfluidic scale effect. To the best of our knowledge, this is the first report of shortening a photocatalytic reaction from several hours to a few seconds, reaching the fastest photocatalytic reaction speed reported in the literature.
In addition to high synthetic reaction speeds, the characterization speed of the reaction products or reactants needs to match the reaction speed to achieve real high-throughput screening. Currently, the routine analysis of organic synthesis products is usually performed by chromatography, mass spectrometry and nuclear magnetic resonance spectroscopy (NMR), most of which are difficult to achieve high-throughput and rapid analysis at the second level34. Spectroscopic analysis has the advantages of non-destructive, rapid and easy to implement online detection. Although its weaknesses in qualitative analysis capability and selectivity limit the application for accurate analysis of reaction products, it is well suited in many systems for online monitoring the changes of unsaturated groups in the synthetic reaction process to provide decisive data for the initial high-throughput screening. Before and after the cycloaddition reaction performed in this work, there was a significant change in the UV spectrum of the reaction solution (caused by the consumption of the substrate) in the region of 280 nm–320 nm, thus the reaction solution could be online detected by coupling a capillary flow cell to the LCW photocatalytic microreactor and using a UV-Vis spectrometer with a detection response time of 0.1 s. We also examined the reliability of the UV-Vis detection method in reaction monitoring using the routine characterization methods for organic synthesis reactions—gas chromatography (GC) and NMR.
High-throughput photocatalytic synthesis and screening
In addition to the ability of rapid synthesis and online characterization, a practical high-throughput system must have the ability to automatically change both continuous variables (e.g., reaction temperature, time, pressure, light intensity, and reactant concentration, etc.) and discrete variables (e.g., substrate, photocatalyst and solvent species, etc.) for performing meaningful high-throughput screening of synthesis conditions. In many cases, much attentions are usually paid to achieving rapid organic synthesis reactions under specific conditions, the ability to achieve rapid screening of a large number of different conditions is frequently overlooked. In the present system, we used a liquid handling module consisted of a selective valve and a syringe pump to achieve automated multistep liquid handling operations, including rapid introducing, changing, mixing and driving of different reaction solutions with adjustable flow rates. We uniquely used a combination of a 1-mL and a 5-mL syringes to perform multi-step reciprocal aspirating-dispensing operations for rapidly achieving automated mixing of mL-scale of reactants (substrate, photocatalyst and solvent) within 2 min. We designed a LabVIEW-based program to control the operation of the whole modules of the system (Fig. 1C), allowing the on-demand automated changing of photocatalyst and substrate species, concentration, photocatalyst ratio, laser light intensity and flow rate to achieve the fully-automated operations of the large-scale screening of reaction conditions (>10,000 reactions) without the need of any manual operation and intervention. This photocatalytic synthesis and screening system also served as one of multiple functional islands of the iChemFoundry (IC) platform. The IC platform is a large-scale automated platform for molecular manufacturing which consisted of various functional islands for realizing a series of automated operations of chemical synthesis from reactant preparation, pre-treatment, chemical synthesis reactions, post-treatment, characterization and data analysis. By combining the functional islands of the photocatalytic synthesis and screening system, the reactant preparation system, an orbital robot and an on-island robotic arm of the IC platform, complete automation of all operations could be achieved, including solid reagent weighing, liquid reagent metering, preparation of reagent stock solutions, transferring and aspirating of reagent stocks, preparation and introduction of reactant solutions, on-line flow photocatalytic reactions, UV spectroscopic reaction monitoring, as well as cyclic screening of different conditions (Fig. 1D).
With the present system, we performed a comprehensive screening for the photocatalytic [2 + 2] cycloaddition reaction with a total number of up to 12,000 conditions, including two discrete variables of the photocatalyst and substrate species and four continuous variables as the laser light intensity, concentration, flow rate, and photocatalyst ratios (Fig. 2). Each of the six variables had 4–6 levels or species, resulting in a total orthogonal combination number of 12,000 screening conditions (i.e., photocatalytic synthesis and analysis experiments). In our system, the average time required to obtain the data for each photocatalytic [2 + 2] cycloaddition reaction was only 32 s, which enabled the system to reach a screening throughput of 2600 conditions per day.
AI-assisted ultra-high-throughput photocatalytic synthesis and screening
Although the above throughput of 2600 conditions per day with a single flow-reactor format was higher than the highest throughputs (e.g., 1500 reactions per day) in the automated synthesis and screening systems reported so far35–37, we noticed that it had the potential to be improved further.
Steady-state and non-steady-state experimental mode
During the high-throughput screening experiments for different reaction conditions, when the synthesis and characterization of the previous condition experiment were completed in the flow system, a new reaction solution needed to be introduced into the reactor and detection flow-cell channels. Due to the convection and molecular diffusion effects existed in the flow system, the previous reacted solution and the newly-introduced unreacted solution would mix with each other at their junction region, and the absorbance signals detected by the UV-Vis detector exhibited a dynamically-changing format (Fig. 3A1) during the switching process of different solutions. The conventional method is to wait for the newly-introduced unreacted reaction solution to flush all of the previous reacted solution out of the flow system to enable the detector to obtain a stable plateau-type steady-state absorbance signal for data reading. However, such a steady-state experimental mode requires much longer waiting time for obtaining the steady-state signals. In the above photocatalytic screening experiment under the steady-state mode, the lasers kept irradiating the microreactor channel and the system spent most of the time (27 s of the 32 s of one experimental cycle time) in switching the different experimental conditions and waiting for a steady-state detection signal to be obtained (for example, as shown in Fig. 3A2). Such a waiting time far exceeded the actual time (<4 s) for photocatalytic synthesis and characterization for a reaction solution, severely limiting the screening throughput for different reaction conditions. In fact, this is one of the major limiting bottlenecks in the application of current flow chemistry systems to high-throughput screening.
To increase the efficiency of time utilization and screening throughput, we proposed the strategy of non-steady-state experimental mode instead of the steady-state mode by using the laser pulse irradiation method to turn the irradiation laser on and off for achieving the rapid switching between the reacted and unreacted solutions, producing a series of non-steady-state continuous peak-shaped signals as shown in Figure 3A3. Under the non-steady-state mode, the waiting time for reaction solution switching was shortened to 6 s and the average time for each experimental cycle was shortened to 8.5 s, achieving an ultra-high throughput up to 10,000 reaction conditions per day (Fig. 3A3, A4, B).
However, these non-steady-state peak signals included the combined absorbance information of the reactants and products from the previous reacted and the newly-introduced unreacted reaction solutions, which were influenced by multiple factors related to the convection and molecular diffusion effects, such as the reaction solution flow rate, the inner diameters and lengths of the reactor and the detection flow-cell channels, and the reactants and products molecular weights. Therefore, it was a great challenge to acquire the corresponding steady-state absorbance data from the non-steady-state peak signals for evaluating the reaction progress.
AI-assisted absorbance prediction
For achieving this complex and challenging task, we developed the AI-assisted absorbance prediction method by using the AI method to analyze the influencing factors related to the convection and molecular diffusion effects and decoupling the non-steady-state data of the adjacent reaction solutions mixed with each other, to predict the corresponding steady-state absorbance data of the respective reaction solutions.
In order to obtain accurate prediction results, we attempted to use 10 regression models based on the principles of linear models, decision tree, neural networks and integrated learning, to process the large numbers of the non-steady-state absorbance data and to predict the corresponding steady-state absorbance data under the same reaction condition, from which we searched for the best-performing model. In these models, the non-steady-state absorbance data (i.e., 40 absorbance data points recorded for each non-steady-state signal peak, Fig. 3A4) as well as the all corresponding 8 variables (i.e., flow rate, laser light intensity, wavelength, substrate concentration, photocatalyst ratio, photocatalyst concentration, substrate species, and photocatalyst species) of the present flow photocatalytic system were set as the inputs of the models (Fig. 4A). The corresponding steady-state absorbance data obtained experimentally using the steady-state mode with the same conditions (i.e., real steady-state absorbance data) were set as the targets of the models, and the predicted steady-state absorbance data were set as the outputs of the models. The substrate and photocatalyst species were input to the models in the form of relative molecular weights instead of chemical structures, since they are directly related to the molecular diffusion effect. On the basis of the massive amounts of the output and target data of the 12,000 reaction conditions (Fig. 4B1), we evaluated the performance of the 10 regression models based on the R2 and RMSE values of the test set. The RMSE values were calculated based on the targets (i.e., real steady-state absorbance data) and the outputs (i.e., predicted steady-state absorbance data) of the models. Among these models, the XGB regression model demonstrated the best prediction performance (Fig. 4B2), which had the smallest RMSE of 0.0140 and the largest R2 of 0.991 with the 70: 30 proportion of the training and test set. We tried to use less data to predict more data (such as 2.5% of the data as the training set and 97.5% of the data as the test set) to further test the predictive performance of the XGB regression model. Pretty good result was still obtained where 300 conditions data were used to predict the remaining 11,700 conditions with a RMSE of 0.0550 and R2 of 0.859 (Fig. 4B3).
These results showed that with the use of the non-steady-state mode and the AI-assisted absorbance prediction method, the long-standing challenge limiting the improvement of screening throughput of flow chemical screening systems caused by inefficient and time-consuming condition switching could be solved. Correspondingly, the screening throughput for the photocatalytic [2 + 2] cycloaddition reaction conditions increased from 2600 to 10,000 conditions per day using the non-steady-state experimental mode, which is the highest level reported in the field of organic synthesis so far. If the manual mode of conventional batch photocatalytic organic synthesis experiments is used to complete the same workload, it would require 2000 organic synthesizers to work one day or a synthesizer to work 2000 days, assuming that each person could complete 5 condition screening experiments per day. What’s more, we repeated all the non-steady-state experiments 3 times to ensure the repeatability, and all 36,000 peak signals are shown in Fig. 3. In addition, in terms of reagent consumption, only 4.0 mmol of each substrate and 0.05 mmol of each photocatalyst were required to complete the whole screening experiment.
We used 10% of the non-steady-state absorbance data as the training set of the XGB model to predict 12,000 steady-state absorbance data, and a 10-fold cross-validation study was performed and the results are shown in Supplementary Information (Table S9). All the absorbance data were converted into product yields (Fig. 4C1). There are 25 data squares in the heatmap, each of which consists of the data of 1 substrate and 1 photocatalyst condition with different concentrations, photocatalyst ratios, flow rates, and laser light intensities (Fig. 4C2). These data squares show different profiles of the product yield distribution, indicating that the photocatalyst and substrate species have significant impact on the product yields. Among the four variables, although all of them show effects on the product yields, the effects of the flow rate and laser light intensity are more obvious. Therefore, we further present the 12,000 data using multidimensional bubble plots with the x-axis of the flow rate and the y-axis of laser light intensity.
Factors influencing photocatalytic [2 + 2] cycloaddition reaction
In the large majority of experiments, the yield of the reaction products significantly increased with the increase of the laser light intensity (Fig. 5A), indicating that sufficient high light intensity is a necessary prerequisite for the high-efficient photocatalytic [2 + 2] cycloaddition reaction in the LCW photocatalytic microreactor.
As shown in Fig. 5A, the product yield increased as the flow rate of the reaction mixture decreased. A higher flow rate means a higher throughput while a shorter residence time (i.e., light irradiation time) of the reaction solution in the photocatalytic microreactor, which results in a lower reaction yield (Fig. 5A). The present high-efficiency LCW photocatalytic microreactor could overcome the contradiction between the throughput and product yield, with which the photocatalytic reaction speed could be dramatically increased compared to the batch synthesis methods38–41, while a high level of product yield and selectivity could also be ensured. For example, in the experiment for S-1, a residence time of 3.3 s in the photocatalytic microreactor (corresponding to the flow rate of 200 μL/min) was sufficient to ensure that the substrate S-1 was completely and selectively converted to the target product.
As shown in Fig. 5B, the optimal photocatalyst for each of the five substrates was quite distinct, which reflects the significance of the high-throughput screening. For S-1, S-4 and S-5, Ir-1 was the optimal photocatalyst with the highest yields, while the photocatalyst Ir-5 was the optimal photocatalyst for S-2 and the photocatalyst Ir-3 was the optimal photocatalyst for S-3.
For the effect of the photocatalyst ratio, generally, the higher the photocatalyst ratio, the faster the reaction speed and the higher product yields (Fig. 5A). For meeting the requirements of green chemistry, 1 mol% photocatalyst is an optimal cost-efficiency choice under a sufficient light intensity.
For the effect of the substrate concentration, in most of the plots (Fig. 5C), we did not observe an evident correlation between the product yields and the substrate concentrations of S-1, S-2, S-3, S-4, and S-5 in the tested range of 0.005 M to 0.02 M, which are frequently adopted in conventional photocatalytic reaction experiments. The reason for this phenomenon may be that our system has an ultra-high catalytic capacity which did not show a significant difference in product yield in the tested range of the substrate concentrations.
In the conventional batch photocatalytic systems, how to increase the substrate concentrations while ensuring sufficiently high yields has been a critical challenge and has not been well addressed so far. For example, in the photocatalytic [2 + 2] cycloaddition reaction reported by Yoon et al. using a conventional batch photocatalytic reactor, when the substrate concentration increased from 0.01 M to 0.05 M, the product yield reduced from 89% to only 33%30. We further increased the concentration of substrate S-1 using 50 μL/min flow rate (i.e., 13.2 s residence time) in the present system, the results showed that the product yield decreased slightly from 91% to 85% when the substrate S-1 concentration increased from 0.01 M to 1.0 M, and further reduced to 67% when the substrate S-1 concentration reached its solubility limit of 2.0 M (Fig. 5D). Compared with that (0.01 M) of the conventional batch system, the substrate concentration could have a 100-fold increase without evidently affecting the reaction yield. We further reduced the flow rate of the 2.0 M substrate reaction solution, the yield could reach 89% at 15 μL/min (i.e., 44 s residence time) (Fig. 5E). The high yields obtained at high substrate concentrations could be attributed to the high light intensity and the fast mass transfer effect in the present photocatalytic microreactor. Such a breakthrough result is of great significance to the application of the photocatalytic cycloaddition reaction in the industrial production of related drugs.
AI-assisted cross-species prediction
To further utilize the above 12,000 data and preliminarily explore the potential possibility of applying AI technique to intelligent chemical synthesis screening, we used the XGB algorithm to perform AI-assisted prediction of product yields cross-substrate and cross-photocatalyst. For cross-species prediction of product yields, the inputs of the models were the reaction conditions (i.e., the 8 reaction variables), the targets were the product yields obtained experimentally, and the outputs were the predicted product yields (Fig. 6A). Differing from the AI-assisted absorbance prediction, the cross-species prediction required detailed chemical structure information of the substrate and photocatalyst species, which were described, digitized, and input to the models using the SMILES strings to generate Modred descriptors (Table S1). The Mordred descriptor dimensions of each substrate and photocatalyst species were reduced to 2 in order to match the dimensions of the other variables and facilitate visualization in subsequent study, as described in Methods. The inputs consisted of 10 dimensions, with 2 representing the substrate species, 2 representing the photocatalyst species, and the remaining 6 representing the other variables (Fig. 6A).
Both the results of the cross-species prediction with different training set ratios and the cross-validation studies are described in the Supplementary Information (Fig. S11, Table S10, S11). The 12,000 experimental data were divided into training and test sets with different ratios for AI-assisted cross-species prediction. As a typical result of the cross-substrate prediction, we used the data of S-1, S-2, and S-4 as the training set to predict the yields of S-5, achieving MAE = 0.0698 and RMSE = 0.0878 (Fig. 6B1). The distinct effects of the 5 photocatalyst species on S-5 were accurately predicted and Ir-1 was predicted as the optimal photocatalyst, which is consistent with the experimental results (Fig. 6B2). With a smaller training set of S-1 and S-2, the prediction for S-4 and S-5 achieved with MAE = 0.0772 and RMSE = 0.0999 (Fig. 6B3, B4). For the cross-photocatalyst prediction, the data of Ir-1, Ir-2, Ir-4 and Ir-5 could be used to predict the yields of Ir-3, with MAE = 0.0364 and RMSE = 0.0497, which presented similar results to the real product yields (Fig. 6C1, C2). When the training set was reduced to include three photocatalysts of Ir-2, Ir-4 and Ir-5, pretty good prediction for Ir-1 and Ir-3 could still be obtained with MAE = 0.0667 and RMSE = 0.0898 (Fig. 6C3, C4). These surprising results exemplarily demonstrated the attractive possibility of utilizing AI algorithms for cross-substrate and cross-photocatalyst prediction of organic synthesis reactions.
Discussion
In summary, we developed a fully-automated and integrated system for high-throughput chemical synthesis, online characterization, and large-scale of condition screening of photocatalytic reactions. With the LCW, microfluidic liquid-handling, and AI techniques, the present system achieved fast photocatalytic cycloaddition reaction in only a few seconds, and reached a high-throughput screening of 10,000 reaction conditions per day that has never been reached before. It can produce massive amount of high-quality data with extremely-low time and reagent consumption, which is difficult to accomplish using conventional systems, providing a powerful platform to facilitate the application of AI techniques in chemical synthesis and to help researchers explore the unknown chemical space more efficiently. As we initially demonstrated in AI-assisted cross-species prediction, such a large amount of data from the same experimental system could provide a solid data base for AI applications. In the future, it would be meaningful to make full use of the 12,000 data and further incorporate AI techniques, such as Bayesian optimization method, for rapid optimization of new products.
In addition, the system can be further improved to meet the needs of different types of photocatalytic reactions, for instance, changing the light source with other wavelengths or trying other types of reactions such as photocatalytic cross-coupling reactions to extend its application scope. For the scale-up of photocatalytic reactions for industrial production, the present system showed a unique potential for increasing the concentration of photocatalytic substrates. In addition to organic synthesis, the application of the system can also be extended to other synthesis fields such as photoinduced material synthesis or biological molecule synthesis in the future.
Methods
Building of the robotic system
The robotic system consisted of a LCW photocatalytic microreactor module, a liquid handling module, and an online UV-Vis absorption spectroscopy module, which is capable of achieving rapid synthesis and high-throughput condition screening of photocatalytic synthesis reactions. We designed a new photocatalytic microreactor based on the LCW principle, which could significantly increase the irradiation light intensity inside the flow microreactor and realize rapid photocatalytic synthesis. In the LCW photocatalytic microreactor module, light from 4 lasers was coupled into flow microreactor with 4 optical fibers to provide high and uniform light irradiation, while the reaction mixture temperature was under control of the cannula-shaped circulating-water temperature control device in order to ensure the selectivity of target products. The inlet and outlet of the LCW photocatalytic microreactor module were connected to the liquid handling module and the online UV-Vis absorption spectroscopy module by capillaries, respectively, allowing the reaction mixture to flow through. The liquid handling module utilized a syringe pump and a 10-port selective valve for liquid handling in the system, including automated introducing of different reactants, preparing reaction mixtures and delivering them through the LCW photocatalytic microreactor and UV-Vis detection module at set flow rates. In order to realize high-throughput detection that could match the efficiency of high-throughput synthesis, the absorption spectra of the reaction mixture after synthesis were recorded in real time by the online UV-Vis absorption spectroscopy module. The operation of all the modules of the robotic system was automatically controlled by a self-written program without human intervention.
LCW photocatalytic microreactor module
In order to realize the LCW-based light illumination and temperature control of the reaction mixture, the LCW photocatalytic microreactor was designed as a cannula configuration with an inner fused-silica capillary (530 μm i.d. and 690 μm o.d., Refined Chromatography Co., Yongnian, China) and an outer glass capillary (1.5 mm i.d. and 2.5 mm o.d., Jingke, Guangzhou, China). The inner fused-silica capillary served as the reaction channel with the reaction mixture solution flowing through, and the outer glass capillary acted as the temperature control channel with the circulating cooling solution flowing through (Fig. S1A). The inner and the outer capillary were installed with two polypropylene tees (3.2 mm, 1/8, Jieliante Technology Co., Beijing, China). The inlet and outlet of the reaction channel were connected with a PFA capillary (1512 L, IDEX, Northbrook, USA) and the detection capillary (250 μm i.d. and 375 μm o.d., Refined Chromatography Co., Yongnian, China), respectively, using unions (P-702, IDEX, Northbrook, USA) or tees (P-712, IDEX, Northbrook, USA).
To provide sufficiently high laser light intensity, the light source consisted of 4 lasers (5.5 W, λmax = 450 nm, Oxlasers, Shanghai, China), from which 4 laser beams were coupled and conducted in the LCW photocatalytic microreactor through 4 optical fibers (400-μm diameter, Lanwin Technology, Zhongshan, China), respectively (Fig. S1B). Each laser beam with a focused spot diameter of ca. 400 μm could be fine-tuned by an x-y-z translation stage to couple as much light as possible into the optical fiber. The 1-meter-long optical fibers consisted of a core/cladding/coating three-layer structure made of quartz/quartz/acrylic with a diameter of 400/440/700 μm and a refractive index of 1.446/1.43/1.51, respectively. With one end of the optical fiber connected with the laser for optical coupling, the other end of the optical fiber was placed between the inner and outer channels to introduce the laser light, and there were two optical fibers both on the left and right sides of the LCW photocatalytic microreactor for providing uniform light illumination (Fig. S1C). The light emitted from the optical fibers spread out naturally according to the numerical aperture (NA) of 0.22, and could be uniformly distributed in the microreactor by multiple total reflections, based on the LCW principle. With the aid of the TracePro software (7.4.3, Lambda Research Co., Littleton, USA), we simulated the light intensity distribution in the cross-section of the reaction channel of the LCW photocatalytic microreactor, which showed that the main section of the reaction channel was under the illumination of strong and uniform laser light (Fig. S1D).
In order to deflect the unwanted heat caused by the high-intensity laser light and to perform the synthesis at set temperature, the cooling solution was circulated in the outer capillary using a diaphragm pump (KVP004, Kamoer Fluid Technology Co., Shanghai, China), and the temperature of the circulating cooling solution was monitored in real time by a digital temperature controller (XH-W3002, Xinghe Electronic Technology Co., Suqian, China). The on and off of the semiconductor chilling plate (TEC1-12706, Xinghe Electronic Technology Co., Suqian, China) was controlled by the digital temperature controller to maintain the temperature of the reaction mixture (Fig. S1E). When the temperature of the cooling solution was set at 20–21 °C, the actual temperature of the reaction mixture in the reaction channel could be maintained in the range of 22.4–27.2 °C measured by an ultrafine thermocouple.
Liquid handling module
In the liquid handling module, the reaction mixture was prepared using a syringe pump (20738325 Cavro XCalibur Pump, Tecan, Mannedorf, Switzerland) and a 10-port selective valve (C5-3000EUTA, VICI, Schenkkan, Switzerland). In order to ensure the data quality of the condition screening, the stock solutions of different reactants need to be freshly prepared every day, which were made in 20 mL vials as 0.2 M substrates and 0.0035 M photocatalysts in DMSO. For performing automated liquid handling, 7 positions of the 10-port selective valve were utilized to select the stock solutions of different reactants for condition screening, with position 4 connected to the solvent, position 3 connected to the substrate, and the remaining 5 positions (i.e., position 1, 2, 5, 6, 7) connected to different photocatalyst species. For the syringe pump, the valve position 1 was connected to the 10-port selective valve for introducing the reactants, and the valve position 2 was connected to a 5 mL syringe that worked in conjunction with another 1 mL syringe for automated dispensing and rapid mixing (Fig. S2A). With the innovative use of these two syringes of different volumes, the reactants could be mixed thoroughly by quickly aspirating and mixing at a rate of more than 50 mL/min for several times according to the mixture volume. Subsequently, the reaction mixture was introduced into the reaction channel of the LCW photocatalytic microreactor through valve position 3 of the syringe pump.
During the reagent sampling process, the solvent DMSO was sampled in the first step and was used to rinse the reaction channel in the last step. The stock solutions of the substrates and photocatalysts were sampled in the intermediate steps to facilitate the homogeneous mixing of the reaction mixture. In order to save time in preparation and ensure high throughput, the volume of reaction mixture prepared each time should be sufficient to complete the screening for a set of 30 conditions with different flow rates and laser light intensities (5 × 6 = 30). In the screening for each set of 30 steady-state conditions, at least 15 reciprocal aspirating-dispensing operations were required to achieve the complete mixing of 5 mL of reaction mixture solution, which took 2 min 30 s. In the screening for each set of 30 non-steady-state conditions, at least 10 reciprocal aspiration-dispensing operations were required to achieve the complete mixing of 3 mL of reaction mixture solution, which took 1 min 15 s. After screening conditions of 6 different laser light intensities at the same flow rate, the reaction channel was rinsed with the unreacted mixture solution to allow the absorbance signal back to the baseline before changing the flow rate. When changes in substrate concentration, photocatalyst ratio, photocatalyst or substrate species were required, it was necessary to thoroughly clean the reaction channel and the detection flow-cell channel with solvent DMSO to avoid cross-contamination. In addition, we tested the performance of the system by alternatively performing condition screenings on reaction mixture of S-1 (0.01 M) catalyzed by Ir-1 (1 mol%) and reaction mixture of S-1 (0.01 M) catalyzed by Ir-4 (1 mol%), using operations of changing the stock solutions with the 10-port selective valve. Each set of experiments was repeated 3 times. The results showed good reproducibility with RSD below 5% (Fig. S2B).
Online UV-Vis absorption spectroscopy module
In the online UV-Vis absorption spectroscopy module, a 6 cm-long fused-silica capillary (250 μm i.d. and 375 μm o.d., Refined Chromatography Co., Yongnian, China) was connected to the LCW photocatalytic microreactor module as a detection flow-cell with an effective detection path length of 240 μm, and the protective layer was removed to allow UV-Vis light to pass through. UV-Vis spectra were recorded on a spectrometer (QEPro, Ocean Optics Co., Dunedin, USA) using a light resource (250-2500 nm, DT-MINI2-GS, Ocean Optics Co., Dunedin, USA) and 2 optical fibers placed on both sides of the detection capillary for light output (OP600-2-UV-VIS, Ocean Optics Co., Dunedin, USA) and acquisition of the absorbance signal (OOS-003946-13, Ocean Optics Co., Dunedin, USA), respectively.
In order to verify the feasibility of this online UV-Vis absorption spectroscopy method, the absorption spectra of carmine pigment as a standard sample were compared in a conventional 10 mm cuvette and in the present 250 μm detection capillary, and the result showed a good fit between the two ones at the wavelengths longer than 260 nm (Fig. S3A). However, when the wavelength was less than 260 nm, the absorbance measured in the detection capillary was slightly shifted compared to those measured in the 10 mm cuvette, which may be related to the relatively weak intensity of the light source in the deep UV wavelength range. Considering that most of the organic synthesis mixture solutions have significant absorption at the wavelengths longer than 260 nm, it is feasible to use this kind of capillary as a detection cell for online UV-Vis absorption spectroscopy detection.
Subsequently, we measured the absorption spectra of each component of the photocatalytic [2 + 2] cycloaddition reaction with the detection module, and there was obvious absorption change of the substrate compared to the products (Fig. S3B). Taking substrate S-1 with the corresponding products P-1 as an example, the products had almost no absorption at the measured wavelength of 293.3 nm. Since the amount of photocatalysts would not consume as the reaction proceeds, theoretically the absorption of the photocatalyst should be constant (Fig. S3C). Catalyzed by photocatalyst Ir-1, the absorbance of the reaction mixture at 293.3 nm gradually decreased as the flow rate decreased, which could indicate the consumption of substrate S-1 and the generation of the products P-1 (Fig. S3D). In the actual experiments, all the UV-Vis absorption spectra information at the wavelength range of 200–1100 nm was recorded to retain as much information as possible, and all absorption spectra were smoothed five times to ensure data quality. For the S-2, S-3, S-4 and S-5 systems, the appropriate wavelengths for absorbance observation were also selected, with 293.3 nm for S-2 and S-5, 285.4 nm for S-3 and 310.7 nm for S-4, respectively.
Basic operation and synthesis procedure of the robotic system
Based on the conditions to be screened, an automated program was written with LabView (8.0, National Instruments, Austin, USA) to control the operation of the 10-port selective valve, syringe pump, adjustable power supply for controlling the laser light intensity and the operation of the on-line UV-Vis spectrometer. In the liquid handling module, different substrates and photocatalyst species were selected, and the desired volumes of the stock solutions were aspirated using the 10-port selective valve and the syringe pump to prepare the reaction mixtures according to the set conditions. After fast and homogeneous mixing, the reaction mixture was delivered into the reaction channel of the LCW photocatalytic microreactor module for rapid synthesis. Unless otherwise mentioned, the photocatalytic [2 + 2] cycloaddition reaction was performed in the LCW photocatalytic microreactor at ambient temperature, with automatic control of the reaction temperature (e.g., 25 ± 2 °C) and uniform light irradiation from the 450 nm lasers. After synthesis, the product yields were evaluated based on the absorbance data obtained by the online UV-Vis absorption spectroscopy module. For the representative conditions, and the GC chromatogram data of the reaction mixture after synthesis was measured with a GC system (7890 A, HP-5 column, Agilent Technologies, Santa Clara, USA) with a flame ionization detector (FID). The detailed procedures followed for each compound and characterization are shown in Supplementary Information (Fig. S8).
AI-assisted prediction of the steady-state absorbance data from the non-steady-state absorbance data
We attempted to use AI technique to process the large numbers of the non-steady-state peak signals and predict the corresponding steady-state absorbance data. First, the non-steady-state absorbance data were aggregated into a large dataset with 12,000 samples, each of which contained 48 eigenvalues (8 reaction variables and 40 absorbance values of a non-steady-state signal peek) and 1 standard output value (steady-state absorbance) (Fig. 4A3).
We tried 3 different principal component analysis (PCA) to downscale the original eigenvalues, and normalized the eigenvalues to improve the model performance. In PCA scheme 1, the 40 non-steady-state absorbance values were downscaled to 1 parameter and then combined with the remaining 8 reaction variables to form new 9 eigenvalues. In PCA scheme 2, we directly downscaled the 48 eigenvalues, while in PCA scheme 3, the 48 eigenvalues were directly used without dimensionality reduction. Taking the XGB model as an example, the model trained with eigenvalues form PCA scheme 1 had the best performance (Fig. S4A).
On the basis of the massive amounts of the data of the 12,000 reaction conditions, using the principles of linear models, decision tree, neural networks and integrated learning, we evaluated 10 regression models including partial least squares regression (PLSR), linear regression, adaptive boosting (AdaBoosting) regression, extremely randomized trees (ExtraTree) regression, support vector regression (SVR), k-nearest neighbor (KNN) regression, multi-layer perceptron (MLP) regression, random forest regression, cascade forest regression, and extreme gradient boosting (XGB) regression (Fig. 3A2). Python (Version 3.6.8) and Microsoft Visual Studio Professional 2017 were used as the development environments. The Python packages including sklearn (Version 0.22), xgboost (Version 1.5.2), and deepforest (Version 0.1.5) were used to train the regression models and to predict the results of the test set. To compare the model performance, the dataset was randomly divided into training and test sets with a proportion of 70% and 30%, respectively. The training set was subjected to k-fold cross-validation, and the hyperparameter combinations of the models were optimized by grid search to improve the model performance. The model performance evaluation metrics RMSE and R2 were calculated using the corresponding built-in functions within sklearn.metrics.
Using the XGB model which showed the best performance among all 10 AI regression models, we respectively used 10% and 30% of non-steady-state data as the training sets to predict the steady-state absorbance data. Then, all the 12,000 absorbance data (including 5 substrate species, 5 photocatalyst species, 4 substrate concentrations, 4 photocatalyst ratios, 5 reaction flow rates, and 6 laser light intensities) were converted to product yields, and the absolute error between the non-steady-state and steady-state product yields are shown in the heatmaps (10% in Fig. S4B, 30% in Fig. S4C).
Descriptor and algorithms used in AI-assisted cross-species prediction
In order to conduct the cross-substrate and cross-photocatalyst prediction, Mordred descriptor was employed to represent different substrates and photocatalysts, which is a popular choice of molecular descriptor calculation software in cheminformatics studies. Mordred descriptor is a combination of known structural and functional properties, such as topological indices and adjacency matrix, and its robustness and comprehensive descriptor calculation capabilities make it well-suited for our prediction task.
With the SMILES string of each molecule as part of the inputs of the models (Table S1), Mordred could generate both 2D and 3D descriptors, encompassing a list of more than 1800 descriptors. However, such a dimension of the Mordred descriptors was very high compared with other numerical variables (i.e., laser light intensity, flow rate, substrate concentration, photocatalyst ratio, photocatalyst concentration, and absorbance wavelength), which may hamper the predictive capabilities of the model. Given the specificity of our problem, where the number of Mordred descriptors far exceeded those of other numerical variables, we did not perform feature selection. Instead, we reduced the dimension of Mordred descriptor of each substrate and photocatalyst species from more than 1800 to 2 with T-distributed stochastic neighbor embedding (TSNE) algorithm, which is a widely used unsupervised dimension reduction technique owing to its advantage in capturing local data characteristics and revealing subtle data structures. Given the large number of Mordred descriptors, the TSNE algorithm is preferable due to its ability to manage high-dimensional data. The reduced Mordred descriptors were then concatenated with other 6 numerical variables to construct the final reaction fingerprints.
We used the XGB algorithm for cross-substrate and cross-photocatalyst prediction, which was a highly efficient and flexible machine learning algorithm based on the gradient boosting framework. It is renowned for its outstanding performance and high efficiency, which are optimized through parallel processing and tree-pruning. Additionally, the XGB’s regularization method could prevent overfitting, thereby improving the model’s generalizability.
Supplementary information
Acknowledgements
The authors acknowledge the financial support provided by the National Ministry of Science and Technology (Grants 2021YFA1301601 to Q.F., 2021YFA1500200 to Z.L. and 2021YFF0701600 to Z.L.), the National Natural Science Foundation of China (Grants 22234007 to Q.F. and 22271249 to Z.L.), the Fundamental Research Funds for the Central Universities (Grants 226-2022-00224 and 226-2023-00115 to Z.L.), the Zhejiang Provincial Natural Science Foundation of China (Grant LDQ24B020001 to Z.L.) and R&D Program of (Grant 2024C03144 to Q.F.), and the iChemFoundry Platform Cooperation Fund of ZJU-Hangzhou Global Scientific and Technological Innovation Center (to Z.L., J.-Z.P., and Q.F.).
Author contributions
J.-M.L. developed the workflow, designed and carried out experiments, analyzed the data and wrote the manuscript. H.-F.W. contributed the idea of photoreactor and rapid mixing method, wrote and developed the control software. J.-W.W. performed AI prediction methods of absorbance data and developed the absorbance data processing software. Q.-H.G. and T.-T.L. synthesized, isolated, and characterized the compounds, and performed NMR analysis. K.-X.C. performed AI algorithms and methods for cross-species prediction. M.-T.Z. helped to perform the light distribution simulations. J.-B.C. constructed the three-dimensional figure of the photoreactor. Q.-N.S., Y.H., and S.-W.S. helped with GC analysis. G.-Y.C. supervised the AI-assisted cross-species prediction. J.-Z.P. helped to develop the control software and supervised the flow manifold design. Z.L. supervised the photocatalysis synthesis work and secured funding. Q.F. conceived the idea, supervised the project, revised the manuscript, and secured funding.
Peer review
Peer review information
Nature Communications thanks Mehrad Ansari and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
Source data are provided with this paper. All data are available from the corresponding author upon request.
Code availability
The source code and dataset are available at GitHub https://github.com/LJM-1997/NCOMMS-24-15301-T and Zenodo 10.5281/zenodo.1373869942.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Jia-Min Lu, Hui-Feng Wang, Qi-Hang Guo.
Contributor Information
Guang-Yong Chen, Email: gychen@zhejianglab.com.
Jian-Zhang Pan, Email: kelvonpan@zju.edu.cn.
Zhan Lu, Email: luzhan@zju.edu.cn.
Qun Fang, Email: fangqun@zju.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-53204-6.
References
- 1.Steiner, S. et al. Organic synthesis in a modular robotic system driven by a chemical programming language. Science363, 144–144 (2019). [DOI] [PubMed] [Google Scholar]
- 2.Elliott, L. D. et al. Batch versus flow photochemistry: a revealing comparison of yield and productivity. Chem Eur. J.20, 15226–15232 (2014). [DOI] [PubMed] [Google Scholar]
- 3.Sagmeister, P. et al. Advanced real-time process analytics for multistep synthesis in continuous flow. Angew. Chem. Int. Ed.60, 8139–8148 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Harenberg, J. H., Weidmann, N., Karaghiosoff, K. & Knochel, P. Continuous flow sodiation of substituted acrylonitriles, alkenyl sulfides and acrylates. Angew. Chem. Int. Ed.60, 731–735 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Douglas, J. J., Sevrin, M. J. & Stephenson, C. R. J. Visible light photocatalysis: applications and new disconnections in the synthesis of pharmaceutical agents. Org. Process Res. Dev.20, 1134–1147 (2016). [Google Scholar]
- 6.Tucker, J. W., Zhang, Y., Jamison, T. F. & Stephenson, C. R. J. Visible-light photoredox catalysis in flow. Angew. Chem. Int. Ed.51, 4144–4147 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Coley, C. W., Abolhasani, M., Lin, H. & Jensen, K. F. Material-efficient microfluidic platform for exploratory studies of visible-light photoredox catalysis. Angew. Chem. Int. Ed.56, 9847–9850 (2017). [DOI] [PubMed] [Google Scholar]
- 8.Nguyen, J. D., Reiss, B., Dai, C. & Stephenson, C. R. J. Batch to flow deoxygenation using visible light photoredox catalysis. Chem. Commun.49, 4352–4354 (2013). [DOI] [PubMed] [Google Scholar]
- 9.Davies, I. W. The digitization of organic synthesis. Nature570, 175–181 (2019). [DOI] [PubMed] [Google Scholar]
- 10.Wilbraham, L., Mehr, S. H. M. & Cronin, L. Digitizing chemistry using the chemical processing unit: from synthesis to discovery. Acc. Chem. Res.54, 253–262 (2021). [DOI] [PubMed] [Google Scholar]
- 11.Raccuglia, P. et al. Machine-learning-assisted materials discovery using failed experiments. Nature533, 73–76 (2016). [DOI] [PubMed] [Google Scholar]
- 12.Chan, H. C. S., Shan, H., Dahoun, T., Vogel, H. & Yuan, S. Advancing drug discovery via artificial intelligence. Trends Pharmacol. Sci.40, 592–604 (2019). [DOI] [PubMed] [Google Scholar]
- 13.Paul, D. et al. Artificial intelligence in drug discovery and development. Drug Discov. Today26, 80–93 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Jimenez-Luna, J., Grisoni, F., Weskamp, N. & Schneider, G. Artificial intelligence in drug discovery: recent advances and future perspectives. Expert Opin. Drug Discov.16, 949–959 (2021). [DOI] [PubMed] [Google Scholar]
- 15.Mehr, S. H. M., Craven, M., Leonov, A. I., Keenan, G. & Cronin, L. A universal system for digitization and automatic execution of the chemical synthesis literature. Science370, 101–108 (2020). [DOI] [PubMed] [Google Scholar]
- 16.Segler, M. H. S., Preuss, M. & Waller, M. P. Planning chemical syntheses with deep neural networks and symbolic AI. Nature555, 604–610 (2018). [DOI] [PubMed] [Google Scholar]
- 17.Sanchez-Lengeling, B. & Aspuru-Guzik, A. Inverse molecular design using machine learning: generative models for matter engineering. Science361, 360–365 (2018). [DOI] [PubMed] [Google Scholar]
- 18.Granda, J. M., Donina, L., Dragone, V., Long, D.-L. & Cronin, L. Controlling an organic synthesis robot with machine learning to search for new reactivity. Nature559, 377-–3381 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Venkatasubramanian, V. The promise of artificial intelligence in chemical engineering: Is it here, finally? Aiche J.65, 466–478 (2019). [Google Scholar]
- 20.Bergman, R. G. & Danheiser, R. L. Reproducibility in chemical research. Angew. Chem. Int. Ed.55, 12548–12549 (2016). [DOI] [PubMed] [Google Scholar]
- 21.Baker, M. Is there a reproducibility crisis? Nature533, 452–454 (2016). [DOI] [PubMed] [Google Scholar]
- 22.Tshitoyan, V. Unsupervised word embeddings capture latent knowledge from materials science literature. Nature571, 95–98 (2019). [DOI] [PubMed] [Google Scholar]
- 23.Ley, S. V., Fitzpatrick, D. E., Ingham, R. J. & Myers, R. M. Organic synthesis: march of the machines. Angew. Chem. Int. Ed.54, 3449–3464 (2015). [DOI] [PubMed] [Google Scholar]
- 24.Trobe, M. & Burke, M. D. The molecular industrial revolution: automated synthesis of small molecules. Angew. Chem. Int. Ed.57, 4192–4214 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Schmink, J. R., Bellomo, A. & Berritt, S. Scientist-led high-throughput experimentation (HTE) and its utility in academia and industry. Aldrichimica Acta46, 71–80 (2013). [Google Scholar]
- 26.Duggan, M. P., McCreedy, T. & Aylott, J. W. A non-invasive analysis method for on-chip spectrophotometric detection using liquid-core waveguiding within a 3D architecture. Analyst128, 1336–1340 (2003). [DOI] [PubMed] [Google Scholar]
- 27.Pan, J.-Z., Yao, B. & Fang, Q. Hand-held photometer based on liquid-core waveguide absorption detection for nanoliter-scale samples. Anal. Chem.82, 3394–3398 (2010). [DOI] [PubMed] [Google Scholar]
- 28.Corcoran, E. B., McMullen, J. P., Levesque, F., Wismer, M. K. & Naber, J. R. Photon equivalents as a parameter for scaling photoredox reactions in flow: translation of photocatalytic C-N cross-coupling from lab scale to multikilogram scale. Angew. Chem. Int. Ed.59, 11964–11968 (2020). [DOI] [PubMed] [Google Scholar]
- 29.Plutschack, M. B., Pieber, B., Gilmore, K. & Seeberger, P. H. The hitchhiker’s guide to flow chemistry (II). Chem. Rev.117, 11796–11893 (2017). [DOI] [PubMed] [Google Scholar]
- 30.Lu, Z. & Yoon, T. P. Visible light photocatalysis of 2+2 styrene cycloadditions by energy transfer. Angew. Chem. Int. Ed.51, 10329–10332 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Du, J., Skubi, K. L., Schultz, D. M. & Yoon, T. P. A dual-catalysis approach to enantioselective 2+2 photocycloadditions using visible light. Science344, 392–396 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Kleinmans, R. et al. Intermolecular [2π+2σ]-photocycloaddition enabled by triplet energy transfer. Nature605, 477–482 (2022). [DOI] [PubMed] [Google Scholar]
- 33.Poplata, S., Troester, A., Zou, Y.-Q. & Bach, T. Recent advances in the synthesis of cyclobutanes by olefin 2+2 photocycloaddition reactions. Chem. Rev.116, 9748–9815 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Sans, V. & Cronin, L. Towards dial-a-molecule by integrating continuous flow, analytics and self-optimisation. Chem. Soc. Rev.45, 2032–2043 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Perera, D. et al. A platform for automated nanomole-scale reaction screening and micromole-scale synthesis in flow. Science359, 429–434 (2018). [DOI] [PubMed] [Google Scholar]
- 36.Gesmundo, N. J. et al. Nanoscale synthesis and affinity ranking. Nature557, 228–232 (2018). [DOI] [PubMed] [Google Scholar]
- 37.Santanilla, A. B. et al. Nanomole-scale high-throughput chemistry for the synthesis of complex molecules. Science347, 49–53 (2015). [DOI] [PubMed] [Google Scholar]
- 38.Lin, S. et al. Mapping the dark space of chemical reactions with extended nanomole synthesis and MALDI-TOF MS. Science361, eaar6236 (2018). [DOI] [PubMed] [Google Scholar]
- 39.González-Esguevillas, M. et al. Rapid optimization of photoredox reactions for continuous-flow systems using microscale batch technology. ACS Cent. Sci.7, 1126–1134 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Ruffoni, A. et al. Practical and regioselective amination of arenes using alkyl amines. Nat. Chem.11, 426–433 (2019). [DOI] [PubMed] [Google Scholar]
- 41.Nicastri, M. C., Lehnherr, D., Lam, Y., DiRocco, D. A. & Rovis, T. Synthesis of sterically hindered primary amines by concurrent tandem photoredox catalysis. J. Am. Chem. Soc.142, 987–998 (2020). [DOI] [PubMed] [Google Scholar]
- 42.Lu, J.-M. et al. Roboticized AI-assisted microfluidic photocatalytic synthesis and screening up to 10,000 reactions per day. Zenodo10.5281/zenodo.13738699 (2024). [DOI] [PMC free article] [PubMed]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
Source data are provided with this paper. All data are available from the corresponding author upon request.
The source code and dataset are available at GitHub https://github.com/LJM-1997/NCOMMS-24-15301-T and Zenodo 10.5281/zenodo.1373869942.