This article provides a comprehensive guide for researchers and drug development professionals facing challenges with failed organic reactions.
This article provides a comprehensive guide for researchers and drug development professionals facing challenges with failed organic reactions. It bridges foundational knowledge of reaction failure mechanisms with cutting-edge methodological approaches, including High-Throughput Experimentation (HTE), Machine Learning (ML) condition prediction, and automated optimization platforms. The content delivers practical troubleshooting frameworks for common experimental pitfalls while introducing advanced validation techniques such as computational reaction simulation and informer library screening. By synthesizing traditional chemical intuition with modern data-driven strategies, this resource aims to significantly reduce optimization time and increase synthetic success rates in complex drug discovery and development workflows.
Why should I move beyond OFAT for reaction optimization?
While the One-Factor-at-a-Time (OFAT) method is straightforwardâchanging one variable while holding others constantâit has major limitations for complex systems. OFAT cannot detect interactions between factors [1]. For example, in your organic reaction, a change in temperature might produce different optimal outcomes depending on the catalyst concentration. OFAT would miss this synergy. Furthermore, OFAT is often inefficient and can lead to suboptimal conclusions [1], potentially causing you to overlook the true best conditions for your failed reaction.
What is a more effective alternative to OFAT?
Design of Experiments (DoE) is a statistically rigorous alternative. Techniques like Taguchi's orthogonal arrays allow you to test multiple factors simultaneously across their different levels (e.g., low, medium, high temperature) in a carefully selected subset of all possible combinations [2] [3]. This method systematically explores the parameter space, revealing both main effects and factor interactions with far fewer experiments than a full factorial approach [4]. A real-world application in optimizing a macrocyclization reaction for OLED device performance used a DoE with only 18 experiments to successfully correlate five different reaction factors with the final device outcome [4].
My reaction has many variables; how can I start with DoE without being overwhelmed?
Begin by identifying the factors (variables) you believe are most influential from your failed experiments. The Taguchi method uses orthogonal arrays specifically to handle an intermediate number of variables (3 to 50) efficiently [2]. For instance, an L18 orthogonal array was used to manage five factors at three levels each, requiring only 18 experimental runs instead of the 243 (3^5) required for a full factorial design [4]. This makes it feasible to get meaningful data on several factors without an unmanageable number of experiments.
How can I make my optimized process more robust?
A key goal of the Taguchi method is robust designâfinding factor levels that make your process less sensitive to uncontrollable "noise" variables (e.g., slight impurities, minor equipment variations) [2]. The method uses a loss function concept, aiming to minimize deviation from the target performance (e.g., yield, purity) and thus reduce costs associated with poor quality and variable outcomes [2].
Potential Cause: Unidentified factor interactions and a narrow operational window found through OFAT screening.
Solution:
Potential Cause: The OFAT-optimized conditions in the lab were not robust to the broader range of variations encountered in a different setting.
Solution:
The table below summarizes the core differences between OFAT and DoE.
| Feature | One-Factor-at-a-Time (OFAT) | Design of Experiments (DoE) |
|---|---|---|
| Basic Principle | Change one variable at a time while holding others constant [1]. | Systematically vary multiple variables simultaneously according to a statistical plan [1]. |
| Detection of Interactions | Cannot detect interactions between factors [1]. | Explicitly identifies and quantifies factor interactions [1]. |
| Experimental Efficiency | Low; can be inefficient and miss optimal conditions [1]. | High; uses structured arrays (e.g., Orthogonal Arrays) to maximize information from minimal runs [2] [4] [3]. |
| Statistical Rigor | Low; results are often qualitative or based on assumed independence. | High; based on statistical principles, allowing for analysis of variance (ANOVA) and significance testing [2]. |
| Best Use Case | Preliminary screening of variables or for very simple systems with no suspected interactions [1]. | Optimizing complex systems with multiple variables and suspected interactions; building robust, scalable processes [1] [4]. |
This protocol is based on a published study that optimized a macrocyclization reaction for OLED performance using a Taguchi orthogonal array and machine learning [4].
1. Define Objective and Performance Measure
2. Determine Design Parameters and Levels
3. Create Orthogonal Array and Execute Experiments
4. Device Fabrication and Performance Evaluation
5. Data Analysis and Model Building
6. Model Validation
The following table details key materials used in the featured OLED macrocyclization study [4].
| Item | Function / Role in the Experiment |
|---|---|
| Dihalotoluene Monomer (1) | The starting material for the Yamamoto coupling macrocyclization reaction. Its structure dictates the formation of the [n]cyclo-meta-phenylene ([n]CMP) products. |
| Ni(cod)â | The nickel catalyst essential for mediating the Ullmann-type coupling reaction that forms the macrocyclic carbon-carbon bonds. |
| Bromochlorotoluene (1b) | A modified monomer used to tweak the reaction kinetics at the oxidative addition step, influencing the distribution of different-sized [n]CMP congeners in the product mixture. |
| DMF (Solvent Component) | A co-solvent used in the reaction mixture. Its percentage was a key factor, believed to influence the reaction at the disproportionation step and thus the product distribution. |
| Ir Emitter (3) | The phosphorescent dopant (likely an Iridium complex) responsible for light emission in the final OLED device. It was mixed with the host [n]CMP material. |
| TPBi (2) | An electron transport material. It was sublimated over the emission layer to form the electron transport layer (ETL) of the OLED device, facilitating electron injection and balancing charge transport. |
Q: What are the common failure modes in IC substrates and how can they be diagnosed?
Failures in integrated circuit (IC) substrates can halt production and impact device reliability. The table below summarizes common failure modes, their causes, and diagnostic methods. [5]
| Failure Mode | Common Causes | Diagnostic Methods |
|---|---|---|
| Delamination | Moisture ingress, CTE mismatch between materials, poor lamination practices. [5] | Visual inspection, Automated Optical Inspection (AOI), microsection analysis for internal layers. [5] |
| Warpage | Asymmetrical stack-ups, CTE mismatches, excessive thermal cycling during assembly. [5] | Flatness measurements, specialized warpage gauges at critical process steps. [5] |
| Conductive Anodic Filament (CAF) | Moisture, electrical bias between conductors, weak resin systems. [5] | Cross-sectional analysis, accelerated environmental testing (high humidity/voltage). [5] |
| Via Cracking | Thermal cycling, inadequate plating, poor via design (e.g., stacked microvias). [5] | Microsection analysis, electrical continuity tests, time-domain reflectometry. [5] |
| Surface & Internal Contamination | Poor cleaning during fabrication, contaminated raw materials, improper handling. [5] | Visual inspection, ion chromatography, target chemical analysis. [5] |
Experimental Protocol for Failure Analysis:
Q: How can functional group incompatibility derail a synthesis and what strategies can prevent this?
Functional group compatibility is the ability of different groups to coexist and participate in intended reactions without interfering. [6] A common example is the incompatibility between highly basic or nucleophilic reagents (like Grignard reagents) and acidic protons present in groups like alcohols, which can lead to deprotonation and side reactions instead of the desired transformation. [7]
Preventive Strategies and Solutions:
Experimental Protocol for Protecting an Alcohol:
Q: Beyond sintering, what unexpected mechanisms can cause catalyst deactivation?
While sintering (particle growth) is a well-known cause of catalyst deactivation, research has identified a novel mechanism: the high-temperature decomposition of nanoparticles into inactive, atomically dispersed single atoms. [9] This process can be more severe and rapid than sintering.
Key Findings on Catalyst Deactivation: A study on Pd/AlâOâ catalysts for methane combustion revealed that stability is strongly dependent on nanoparticle density. Counterintuitively, catalysts with a higher density of nanoparticles were more stable, while sparse catalysts deactivated rapidly due to full decomposition of nanoparticles into single atoms. [9]
Experimental Protocol for Probing Catalyst Stability:
Research Reagent Solutions for Catalyst Studies
| Reagent/Material | Function in Research Context |
|---|---|
| Gamma-Alumina (γ-AlâOâ) | A common high-surface-area support material for anchoring metal catalysts. [9] |
| Colloidal Pd Nanoparticles | Pre-synthesized nanoparticles used to impregnate the support, allowing independent control of particle size and loading. [9] |
| HAADF-STEM | High-resolution microscopy technique to visualize nanoparticle distribution, size, and the presence of single atoms after aging. [9] |
| EXAFS | Spectroscopic technique used to determine the coordination number and local environment of metal atoms, confirming the loss of Pd-Pd bonds. [9] |
FAQ 1: What are the most common numerical errors that can occur in a standard DFT calculation? Numerical errors are often introduced through the choice of integration grids and basis sets. Using a grid that is too coarse is a common pitfall, especially with modern functionals. For example, meta-GGA functionals (like M06 or SCAN) and many B97-based functionals perform poorly on small grids and require much larger ones, such as a (99,590) grid, to deliver reliable energies and forces. Furthermore, the RIJCOSX approximation for accelerating integrals can lead to significant non-zero net forces and errors in individual force components if not properly controlled [10] [11].
FAQ 2: My DFT calculation predicts an incorrect reaction barrier or product distribution. What could be the underlying functional error? This is a classic symptom of delocalization error, also known as self-interaction error, where an electron incorrectly interacts with itself. This error can lead to an underestimation of reaction barriers and a misassignment of electronic states, particularly in systems like transition metal complexes or during processes involving charge transfer. A related issue is the "sd energy imbalance" in transition metals, where DFT provides an unbalanced description of electrons in s versus d orbitals, which is crucial for accurately modeling catalysts [12].
FAQ 3: How can I identify if the error in my calculation is due to the functional itself or the electron density? The theory of Density-Corrected DFT (DC-DFT) provides a framework for this. It separates the total error into a functional-driven error and a density-driven error. A practical method to check is to perform a HF-DFT calculation: take the Hartree-Fock electron density and use it to evaluate the DFT energy. If the HF-DFT result is significantly closer to the correct value, your error is likely density-driven, indicating the self-consistent DFT procedure is producing a poor density [13].
FAQ 4: Why do my calculated lattice parameters differ from experimental values, and how can I improve them? Different exchange-correlation (XC) functionals have systematic biases in predicting lattice parameters. For example, LDA tends to underestimate them, while PBE often overestimates them. To improve accuracy, consider using a more modern functional like PBEsol or vdW-DF-C09, which are designed for solids and show lower mean absolute relative errors [14]. The table below quantifies the typical errors for various functionals.
Table 1: Systematic Errors in Lattice Parameter Predictions for Oxides [14]
| XC Functional | Type | Mean Absolute Relative Error (MARE) | Systematic Trend |
|---|---|---|---|
| LDA | Local Density Approximation | ~2.21% | Underestimation |
| PBE | Generalized Gradient Approximation | ~1.61% | Overestimation |
| PBEsol | GGA for solids | ~0.79% | Near zero |
| vdW-DF-C09 | van der Waals Functional | ~0.97% | Near zero |
FAQ 5: I am getting erratic free energies for my reaction. What low-frequency vibrational issue should I check? Low-frequency vibrational modes can artificially inflate entropy contributions due to their inverse relationship with the entropic correction. Spurious low-frequency modes, which may be due to incomplete optimization or inherent molecular flexibility, can lead to incorrect predictions of reaction barriers or stereochemical outcomes. A recommended correction is to raise all non-transition-state modes below 100 cmâ»Â¹ to 100 cmâ»Â¹ for the purpose of computing the entropic correction [11].
Self-interaction error (SIE) is a fundamental flaw where an electron interacts with itself, skewing results for transition metals crucial in catalysis [12].
Protocol:
Diagram: Identifying and Correcting for Self-Interaction Error
Numerical integration grids are a major source of systematic error, especially for free energy calculations and modern density functionals [11].
Protocol:
dftgrid 3 in TeraChem) for all production calculations [11].DEFGRID3 keyword [10].Table 2: Troubleshooting Numerical Errors in DFT Calculations
| Symptom | Potential Cause | Correction Protocol |
|---|---|---|
| Free energy varies with molecular orientation | Coarse integration grid | Use a (99,590) grid or equivalent; test grid sensitivity [11]. |
| Non-zero net force on a stationary molecule | RIJCOSX approximation or loose DFT grid | Disable RIJCOSX or use DEFGRID3 in ORCA; recompute with tight settings [10]. |
| Large errors with mGGA/SCAN functionals | High grid sensitivity of functional | Mandate use of a dense integration grid [11]. |
| SCF convergence failure | Chaotic SCF behavior | Use hybrid DIIS/ADIIS; apply level shifting (e.g., 0.1 Hartree); tighten integral tolerance to 10â»Â¹â´ [11]. |
High-throughput materials screening requires careful functional selection to minimize systematic property errors [14].
Protocol:
Diagram: Workflow for Managing Systematic Errors in High-Throughput Studies
Table 3: Essential Computational Tools and Protocols for Mitigating DFT Errors
| Tool or Protocol | Function | Use Case Example |
|---|---|---|
| Dense Integration Grid (e.g., 99,590) | Minimizes numerical error in evaluating the exchange-correlation functional. | Essential for obtaining reliable free energies with mGGA functionals like SCAN; prevents energy drift upon molecular rotation [11]. |
| Density-Corrected DFT (DC-DFT) | Separates functional error from density error by using a more accurate density (e.g., from HF). | Diagnosing whether error in anion energetics or reaction barriers stems from the functional approximation or the self-consistent DFT density [13]. |
| Self-Interaction Correction (FLOSIC) | Corrects the spurious self-interaction of electrons in approximate DFT. | Improving the accuracy of ionization energies and electronic structure descriptions in transition-metal catalysts [12]. |
| Cramer-Truhlar Low-Frequency Correction | Corrects entropic contributions from spurious low-frequency vibrations. | Raising sub-100 cmâ»Â¹ vibrational modes to 100 cmâ»Â¹ for stable and accurate free energy calculations [11]. |
| Bayesian Error Estimation | Quantifies uncertainty in DFT predictions by analyzing an ensemble of functionals. | Providing "error bars" on predicted material properties in high-throughput virtual screening [14]. |
| Hoechst 34580 tetrahydrochloride | Hoechst 34580 tetrahydrochloride, MF:C27H33Cl4N7, MW:597.4 g/mol | Chemical Reagent |
| Aphadilactone C | Aphadilactone C|DGAT-1 Inhibitor |
What is the primary advantage of using HRMS for failure analysis? High-Resolution Mass Spectrometry (HRMS) enables the sensitive, simultaneous detection of a wide range of compounds in full-scan mode with high mass accuracy (typically ⤠5 ppm) and resolution [15]. This allows researchers to identify unknown byproducts and transformation products without needing reference standards upfront, which is crucial for diagnosing failed organic reactions [15].
My LC-HRMS analysis isn't detecting expected non-polar byproducts. What could be wrong? The chemical domain of your method might be too narrow. Standard reversed-phase LC-ESI-MS is optimized for polar compounds [15]. If you suspect non-polar byproducts, consider these adjustments:
How can I improve confidence in byproduct identification using HRMS? Confidence is built through a multi-step process [15]:
What quality control measures are essential for reliable non-target screening? Implement a robust Quality Assurance/Quality Control (QA/QC) protocol [15]. This includes:
Possible Causes and Solutions:
| Possible Cause | Diagnostic Steps | Solution |
|---|---|---|
| Sample Preparation Issues | Check procedure blanks for contamination. Review recovery rates of internal standards. | Optimize solid-phase extraction (SPE) protocols or use direct injection if concentration allows. Use solvent mixtures (e.g., methanol, acetonitrile) suitable for a broad compound range [15]. |
| Ion Suppression in ESI | Post-infusion analysis to check for signal suppression in complex matrices. | Improve chromatographic separation to isolate analytes. Dilute the sample or perform a cleaner extraction to reduce matrix effects [15]. |
| Carryover Contamination | Run a solvent blank after a high-concentration sample. | Incorporate rigorous needle and column wash steps in the instrumental method. Increase the wash volume between injections. |
Possible Causes and Solutions:
| Possible Cause | Diagnostic Steps | Solution |
|---|---|---|
| Insufficient Chromatographic Resolution | Check peak shape and resolution in the chromatogram. | Optimize the LC gradient (e.g., a broader gradient from 0-100% organic solvent). Consider using a different column chemistry (e.g., HILIC for very hydrophilic compounds) [15]. |
| Lack of Orthogonal Data | HRMS alone may give the same mass for isomers. | Incorporate NMR spectroscopy to elucidate structural connectivity and confirm isomeric identity. Use ion mobility spectrometry (IMS) to separate ions by their size and shape before mass detection [15]. |
Principle: Use minimal and generic sample processing to avoid losing compounds with diverse physico-chemical properties [15].
Materials:
Procedure:
Principle: A structured, step-wise approach is necessary to move from unknown peaks to confident identifications [15].
Procedure:
| Reagent / Material | Function in Failure Analysis |
|---|---|
| Mixed-mode SPE Cartridges | Broadens the range of extractable compounds from a sample by utilizing multiple chemical interactions (e.g., ion-exchange, reversed-phase), crucial for capturing unknown byproducts with diverse properties [15]. |
| Deuterated Solvents (e.g., DMSO-dâ, CDClâ) | Essential for NMR spectroscopy, allowing for the dissolution of reaction mixtures and structural elucidation of byproducts without interfering proton signals. |
| Internal Standard Mixture | A set of isotopically labeled or otherwise unique compounds added to the sample to monitor and correct for variations in sample preparation, matrix effects, and instrument performance [15]. |
| HRMS Mass Calibrant | A standard solution used to calibrate the mass spectrometer, ensuring the high mass accuracy (⤠5 ppm) required for reliable molecular formula assignment of unknowns [15]. |
| Reverse-Phase (e.g., C18) LC Columns | Provides the primary separation mechanism in LC-HRMS, separating compounds based on hydrophobicity with a generic gradient (e.g., 0-100% methanol) to cover a wide chemical range [15]. |
| Aspartocin D | Aspartocin D, MF:C57H89N13O20, MW:1276.4 g/mol |
| Hedycoronen A | Hedycoronen A, MF:C21H30O3, MW:330.5 g/mol |
Q1: Why did my organic reaction fail to produce any product, and how can DOE help?
Failed organic reactions, particularly those yielding no product, often result from errors in initial experimental conditions rather than fundamental chemical theory. Common causes include calculation errors (e.g., misplaced decimal points), improper measurement of reactants, incorrect heating temperatures, or using the wrong reagents (e.g., confusing acetic anhydride for acetic acid) [16]. DOE helps by systematically testing the key factors that influence the reaction outcome. Instead of the inefficient "one factor at a time" (OFAT) approach, DOE allows you to simultaneously investigate multiple variables and their interactions to quickly identify which specific factor or combination of factors is causing the failure [17].
Q2: My experiment worked for my colleague but failed for me. What could be the reason?
This is a classic sign that the failure is due to specific procedural choices rather than a flawed experimental design [16]. Minor misunderstandings in interpreting the lab manual, subtle variations in technique during the work-up and purification phase (such as confusing the phases during liquid-liquid extraction), or minor deviations in handling can lead to drastically different outcomes. A screening DOE is an excellent tool for investigating such operator-dependent variables. It can efficiently narrow down the few critical factors from a long list of potential suspects, helping to pinpoint the exact step where the processes diverge [18].
Q3: How many experimental runs do I need to perform to be statistically confident in my results?
The number of required runs depends on the failure rate you are investigating and the number of factors you wish to study. As a rule of thumb, to validate a solution for a problem with a failure rate of p, you should test at least n = 3/p units and observe zero failures to have statistical confidence (α=0.05) that you have truly improved the process [19]. For example, to address a 10% failure rate, you should plan to test 30 units. Furthermore, the number of experimental runs in a factorial design can be calculated using the formula 2^n, where n is the number of factors [17]. Fractional factorial designs can drastically reduce this number while still providing valuable insights [20] [21].
Q4: What is the difference between common cause and assignable cause variation, and how does DOE address them?
While control charts are used to identify assignable causes, DOE is particularly powerful for tackling common cause variation. It allows you to actively experiment with the process factors (e.g., people, methods, materials, machines) to discover which ones, when changed, can reduce the underlying variability and improve the overall process capability [20].
Q5: How can I efficiently explore a large parameter space with many factors?
When dealing with a high number of factors, a two-stage approach is highly efficient:
This guide outlines a structured methodology, based on the Plan-Do-Check-Act (PDCA) cycle, for diagnosing and correcting persistent experimental failures [20].
Workflow Overview:
Step-by-Step Protocol:
Plan: Define the Problem and Propose a Hypothesis.
Do: Execute a Screening Design of Experiment.
Check: Analyze the Data and Verify the Hypothesis.
Act: Implement and Validate the Solution.
This guide focuses specifically on the critical failure of an organic reaction yielding no product.
Common Failure Points and Diagnostic Flow:
Diagnostic Steps and DOE Applications:
For Suspected Stage 1 Failures:
For Suspected Stage 2 Failures:
| Design Type | Objective | Number of Runs for k=4 factors | Key Advantage | Best Use Case in Organic Chemistry |
|---|---|---|---|---|
| Full Factorial | Study all factors & interactions | 16 (2â´) | Estimates all main effects and interactions | Final optimization when few critical factors are known [17] |
| Fractional Factorial (Res V) | Screen many factors efficiently | 8 (2â´â»Â¹) | Reduces runs while aliasing higher-order interactions | Initial screening to identify critical factors from a large list [21] |
| Definitive Screening (DSD) | Screen factors with curvature | 9 (2*4+1) | Identifies active main effects & curvature in one design | Screening when nonlinear effects (e.g., from temperature) are suspected [20] [22] |
| Response Surface (e.g., CCD) | Model curvature and find optimum | ~25-30 (varies) | Fits a quadratic model for optimization | Finding the ideal temperature and concentration for maximum yield [17] [22] |
| Plackett-Burman | Very efficient screening | 12 for k=11 | Maximum factors for minimal runs | Early-stage scouting of a very large number of potential variables [22] |
| Item | Function in Experiment | Key Consideration for DOE |
|---|---|---|
| Anhydrous Solvents | Provide reaction medium; purity critical for moisture-sensitive reactions | A key factor to test at different grades or water-content levels [16]. |
| Catalysts (e.g., Pd/C, Enzymes) | Accelerate reaction rate and selectivity | Catalyst loading and age/lot are prime candidates as DOE factors. |
| Acid/Base Reagents | Used as reactants, catalysts, or in work-up for pH adjustment | Concentration (e.g., Conc. vs. 6M) and equivalents are critical factors [16]. |
| Purification Media | e.g., Silica gel for chromatography; solvents for recrystallization | The solvent system composition for TLC/column is a common factor to optimize via DOE. |
| Drying Agents | e.g., MgSOâ, NaâSOâ; remove water after aqueous work-up | The type and amount of drying agent can be studied as a factor for optimal yield. |
| Pep1-TGL | Pep1-TGL, MF:C41H71N11O15S, MW:990.1 g/mol | Chemical Reagent |
| Pseudo RACK1 | Pseudo RACK1, MF:C144H225N43O34S3, MW:3198.8 g/mol | Chemical Reagent |
Objective: To investigate the effects of Reaction Temperature (A) and Catalyst Equivalents (B) on the Yield of a model organic reaction.
Methodology:
Define Factor Levels:
Create Design Matrix and Execute Runs:
| Standard Order | Run Order | A: Temp (°C) | B: Catalyst (equiv) | Yield (%) |
|---|---|---|---|---|
| 1 | 3 | 60 (-1) | 1.0 (-1) | 55 |
| 2 | 1 | 80 (+1) | 1.0 (-1) | 70 |
| 3 | 4 | 60 (-1) | 1.5 (+1) | 65 |
| 4 | 2 | 80 (+1) | 1.5 (+1) | 85 |
Calculate Main Effects:
Analysis:
Q1: What is the primary objective of implementing a High-Throughput Experimentation (HTE) workflow? HTE is a process of scientific exploration involving lab automation, effective experimental design, and rapid parallel or serial experiments. Its primary objective is to efficiently navigate high-dimensional design spaces, either for optimization (finding the highest-performing material) or exploration (mapping a structure-property relationship to build predictive models) [25].
Q2: What are the core components of a functional HTE program? A functional HTE program requires three core components [26]:
Q3: My HTE robotic assembly process fails due to part variations. How can this be improved? For complex robotic assembly processes with part variations, online performance optimization methods can be deployed. One effective approach uses Gaussian Process Regression surrogated Bayesian Optimization Algorithm (GPRBOA). This data-driven method constructs a non-parametric model of the assembly process and iteratively optimizes parameters (e.g., insertion force, search speed) to maximize success rate and minimize cycle time without interrupting production [27].
Q4: What are common data analysis challenges in HTE, and how can they be addressed? HTE generates volumes of data that are impossible to process manually. Challenges include data integration, analysis, and leveraging it for decision-making. Success often requires combining Electronic Lab Notebook (ELN) and Lab Information Management System (LIMS) environments. The future of HTE data analysis lies in improved integration with Artificial Intelligence (AI) and Machine Learning (ML) to create reliable predictive models [26].
Q5: A single reaction in my HTE library failed. What are the most likely causes? If a single reaction in an otherwise successful library fails, the issue is typically at the individual reaction level, not the core HTE design. Common causes are calculation errors, improperly measured reactants, use of wrong reagents, or improper heating/reaction times. These errors often occur in the reaction stage rather than the work-up and purification [16].
| Problem Symptom | Possible Root Cause | Corrective Action |
|---|---|---|
| Low Success Rate Across Entire Library | Poorly defined design space; insufficient prior knowledge [25]. | Begin with a small-scale, rationally designed library to troubleshoot synthesis and characterization protocols before full-scale HTE. |
| High Data Volume but Poor Insights | Lack of a FAIR-compliant data environment; inadequate data analysis tools [26]. | Invest in integrated IT/informatics infrastructure (ELN, LIMS) and leverage statistics/ML for data featurization and analysis [25] [26]. |
| HTE Results Do Not Scale Up | Library parallelization/miniaturization is not relevant to production scale, especially in materials science [26]. | Use larger-scale equipment with limited parallelization (e.g., 4-16 reactors) and conditions that allow easier scale-up. |
| Optimization Stuck at Local Maximum | Library design cannot navigate "activity cliffs" where similar materials have very different performances [25]. | Implement adaptive sampling or active learning (AL) techniques to strategically select new experiments and escape local maxima [25] [26]. |
| Problem Symptom | Possible Root Cause | Corrective Action |
|---|---|---|
| Robotic Assembly Failures (e.g., jamming) | Part misalignment due to fixture errors or dimensional variations; sub-optimal assembly parameters [27]. | Implement an online parameter optimization method like GPRBOA to autonomously find optimal parameters (e.g., insertion force, search radius) [27]. |
| Failure to Autonomously Recover from Errors | Lack of real-time failure reasoning and adaptive recovery mechanisms. | Integrate a failure recovery framework that uses Vision-Language Models (VLMs) for real-time failure detection and reasoning, combined with a reactive planner to dynamically correct actions [28]. |
| Inefficient Optimization Cycle Times | Offline optimization algorithms (e.g., Genetic Algorithms) are slow and lack efficiency for online use [27]. | Deploy Gaussian Process Regression with Bayesian Optimization (GPRBOA) for online, iterative optimization that balances exploration and exploitation [27]. |
The following table details key components and their functions in a typical HTE platform, as derived from featured experiments and field overviews.
| Item / Solution | Function / Role in HTE |
|---|---|
| Liquid Handlers & Robotics (e.g., Tecan, Hamilton) | Automate repetitive liquid transfer and synthesis steps, enabling rapid parallel or serial experimentation and reducing human error [26]. |
| Design of Experiments (DOE) Software | Computational tool to strategically design library members and experiments, maximizing information gain while reducing experimental burden [25] [26]. |
| Electronic Lab Notebook (ELN) | Captures experimental protocols, ideation, and raw data in a findable, accessible format, forming the foundation of a FAIR data environment [26]. |
| Lab Information Management System (LIMS) | Manages sample tracking, workflow execution, and integrates with analytical instruments for high-throughput data capture [26]. |
| Gaussian Process Regression (GPR) Model | A non-parametric model used to map the relationship between input parameters (e.g., assembly force) and complex system performance, accounting for uncertainty [27]. |
| Bayesian Optimization Algorithm (BOA) | An optimization strategy that uses a surrogate model (like GPR) to intelligently select the next experiment to run, balancing exploration of new areas and exploitation of known good areas [27]. |
| Vision-Language Models (VLMs) | Provides real-time visual understanding and reasoning for robotic systems, enabling failure detection, identification of root causes, and suggesting corrective actions [28]. |
Problem: Machine learning models fail to accurately predict catalyst performance for reactions outside the training dataset.
Diagnosis Steps:
Solutions:
Experimental Protocol for Data Augmentation:
Problem: Models recommend traditional toxic solvents instead of sustainable alternatives.
Diagnosis Steps:
Solutions:
Bayesian Optimization Protocol for Solvent Selection [33]:
Problem: Models cannot accurately predict interdependent reaction conditions (catalyst-solvent-temperature combinations).
Diagnosis Steps:
Solutions:
Workflow:
Methodology Details:
Workflow:
Performance Metrics:
Workflow:
Implementation Details:
Table 1: Machine Learning Model Performance Metrics
| Model Type | Application | Accuracy/Performance | Dataset Size | Reference |
|---|---|---|---|---|
| Two-Stage Neural Network | Reaction condition prediction | 73% top-10 exact match (solvents/reagents), 89% within ±20°C (temperature) | 74,683 reactions | [30] |
| ANN with Genetic Algorithm | SCR NOx catalyst discovery | Novel Fe-Mn-Ni catalyst identified in 4 iterations | 2,748 data points | [29] |
| Bayesian Optimization | Green solvent selection | Reduced experimental iterations by ~60% | 8 solvent candidates | [33] |
| GPT-4 Assisted Framework | Hydrocracking catalyst optimization | Reduced experimental iterations by 60% | Industrial data | [34] |
| Green Solvent Prediction | Organic solvent recommendation | 85.1% top-3 accuracy, 88% experimental success | Patent-derived reactions | [31] |
Table 2: Data Requirements for Different Model Types
| Model Architecture | Minimum Data Requirements | Optimal Data Size | Key Preprocessing Steps |
|---|---|---|---|
| Artificial Neural Networks (ANN) | ~1,000 data points | 2,500+ points | Feature normalization, outlier removal [29] |
| Two-Stage Neural Network | ~10,000 reactions | 50,000+ reactions | Reaction fingerprinting, chemical standardization [30] |
| Bayesian Optimization | Can start with physical models | 40-100 experiments | Prior knowledge incorporation, uncertainty quantification [33] |
| Random Forest/XGBoost | ~500 samples | 2,000+ samples | Feature selection, hyperparameter tuning [35] |
Table 3: Essential Materials for ML-Guided Reaction Optimization
| Reagent/Material | Function in ML Experiments | Example Application | Key Considerations |
|---|---|---|---|
| Fe-Mn-Ni catalysts | SCR NOx catalyst optimization | Environmental catalysis | Composition control via co-precipitation [29] |
| Green solvent library (8 candidates) | Sustainable solvent screening | Lignin bioproduct separation | Mixture optimization via Bayesian methods [33] |
| NaâCOâ precipitation agent | Catalyst synthesis control | Fe-Mn-Ni catalyst preparation | pH control at 11.5, 17h aging [29] |
| Liquid-handling robot | High-throughput experimentation | Bayesian solvent optimization | 40 simultaneous tests, automation integration [33] |
| Morgan fingerprints (radius 2, 4096 bits) | Reaction representation | Condition prediction | Combines product structure and reactant-product differences [30] |
| Tertiapin LQ | Tertiapin LQ, MF:C106H179N33O24S4, MW:2428.0 g/mol | Chemical Reagent | Bench Chemicals |
| Obtustatin | Obtustatin, MF:C184H284N52O57S8, MW:4393 g/mol | Chemical Reagent | Bench Chemicals |
Q1: Which machine learning model should I start with for predicting reaction conditions?
A: For most applications, begin with a two-stage neural network approach if you have >10,000 reaction examples [30]. For smaller datasets (<2,000 examples), use iterative ANN with genetic algorithm optimization [29]. When exploring completely new chemical spaces with limited data, Bayesian optimization with physical model priors is most effective [33].
Q2: How can I validate ML-predicted reaction conditions before full experimental commitment?
A: Implement three validation tiers: (1) Computational validation using physical models like COSMO-RS for solvent predictions [33], (2) Small-scale (1-5 mL) robotic validation for top candidates [33], (3) Uncertainty quantification to identify high-confidence predictions [31]. This approach reduces failed experiments by 60% [34].
Q3: My model performs well on historical data but fails with new catalyst systems. How can I improve generalizability?
A: This indicates domain shift. Implement three strategies: (1) Add iterative experimental feedback loops to update models with new data [29], (2) Use transfer learning from related catalyst families with abundant data [32], (3) Incorporate physical constraints and rules into ML models to ensure chemically plausible predictions [35].
Q4: What are the minimum data requirements to build a useful prediction model for catalyst selection?
A: Minimum requirements vary by model: ANN models need ~1,000 data points [29], two-stage neural networks require ~10,000 reactions [30], while Bayesian optimization can start with physical model priors and 40-100 experiments [33]. For novel systems, focus on diverse data coverage rather than quantity.
Q5: How can I ensure my model recommends environmentally sustainable solvents?
A: Implement green solvent replacement methodologies that: (1) Incorporate sustainability metrics directly into the loss function [31], (2) Use Bayesian optimization with environmental constraints [33], (3) Maintain performance-based ranking while filtering for green chemistry principles [31]. This achieves 80% success rate for green solvent recommendations [31].
In the pursuit of novel synthetic methods, organic chemists must demonstrate that their new reaction can work on a variety of molecular structures. Traditionally, this is achieved through a substrate scope, which involves running the reaction on a series of different, often specially synthesized, substrates. However, this process can be time-consuming and resource-intensive, sometimes taking six months or more to prepare the necessary complex substrates [36].
Functional Group Robustness Screening has emerged as a powerful, complementary technique to address this bottleneck. Also known as an additive-based screen or intermolecular reaction screening, this method rapidly evaluates the tolerance of a given set of reaction conditions to a wide range of functional groups, as well as the stability of those functional groups to the reaction conditions [37] [38]. By simply adding commercially available compounds containing specific functional groups to the reaction mixture, researchers can gather extensive data on functional group compatibility in a matter of days, rather than months [36].
This guide provides troubleshooting support for researchers implementing these screens in their work on organic reaction development.
Q1: How does robustness screening differ from a traditional substrate scope?
A traditional substrate scope tests the reaction on different, fully-constructed molecules, providing information on steric, electronic, and functional group effects in an intramolecular context. In contrast, a robustness screen tests the reaction's tolerance to foreign, intermolecular additives. It provides two key pieces of information [36]:
The two methods are complementary. The screen quickly identifies potential limitations or points of failure, while the substrate scope confirms the findings in a more synthetically relevant context [36].
Q2: What are the main advantages of using this screening approach?
Q3: What are the limitations of this method?
Q4: My screening results are inconsistent. What should I check?
This protocol, adapted from the literature, describes a general method for evaluating functional group tolerance in a batch format [37].
Step 1: Preparation
Step 2: Reaction Setup
Step 3: Analysis and Data Processing
The following diagram illustrates the key decision points and steps in the robustness screening process.
The table below summarizes example quantitative data from a robustness screen, illustrating how results are typically reported. This data is based on studies evaluating functional group compatibility for various reactions [39] [38].
Table 1: Example Results from a Functional Group Robustness Screen
| Functional Group | Representative Additive | Reaction Yield (%) | Additive Remaining (%) | Tolerance Assessment |
|---|---|---|---|---|
| Aromatic Ring | Toluene | 95 | 99 | Excellent |
| Halide (Aryl) | 4-Bromotoluene | 90 | 98 | Excellent |
| Ether | Anisole | 88 | 97 | Excellent |
| Ester | Methyl Benzoate | 85 | 95 | Good |
| Ketone | Acetophenone | 80 | 92 | Good |
| Nitrile | Benzonitrile | 75 | 90 | Moderate |
| Free Alcohol | Benzyl Alcohol | 45 | 85 | Low (Reaction Inhibited) |
| Free Amine | Pyridine | 20 | 95 | Low (Reaction Inhibited) |
| Aldehyde | Benzaldehyde | 15 | 30 | Low (Additive Degraded) |
| Alkene | Styrene | 10 | 25 | Low (Additive Degraded) |
Table 2: Essential Materials for Robustness Screening
| Item / Reagent | Function / Purpose | Considerations |
|---|---|---|
| Functional Group Evaluation (FGE) Kit | A pre-selected set of additives covering common functional groups (acids, bases, nucleophiles, heterocycles) [39]. | Kits can be assembled in-house from commercial chemicals. Aim for diversity. |
| Palladium on Alumina (Pd/AlâOâ) | A common heterogeneous catalyst used in demonstrative studies, e.g., for nitro group reductions [38]. | Catalyst poisoning can be a specific focus of the screen. |
| Ammonium Salts (e.g., NHâI) | Additives that can act as catalysts or accelerants in certain reactions, such as amide bond cleavage [39]. | Screening can reveal their dual role as reagents and compatibility markers. |
| Hexafluoroisopropanol (HFIP) | A strong hydrogen-bond-donating solvent used to accelerate certain reactions and tested for its effect in screens [39]. | Its unique properties can positively or negatively impact specific functional groups. |
| Gas Chromatography (GC) | A primary analytical technique for rapid quantification of reaction components and additive stability [37]. | Ideal for volatile mixtures; requires method development. |
| High-Performance Liquid Chromatography (HPLC/UHPLC) | An orthogonal analytical technique to GC, used for less volatile or thermally labile compounds [38]. | Provides high-resolution data for complex mixtures. |
| In-line FT-IR Spectroscopy | A Process Analytical Technology (PAT) for real-time reaction monitoring in automated flow systems [38]. | Enables rapid data acquisition; requires advanced data modeling (e.g., PLS). |
| [Pro3]-GIP (Mouse) | [Pro3]-GIP (Mouse), MF:C225H342N62O64S, MW:4972 g/mol | Chemical Reagent |
| Acein | Acein, MF:C43H68N10O13, MW:933.1 g/mol | Chemical Reagent |
For higher-throughput or more detailed analysis, advanced methodologies have been developed.
Integrating functional group robustness screening into your reaction development workflow provides a honest and efficient way to assess the practical utility of a new methodology. To maximize its effectiveness:
Q: My systematic review protocol feels unstructured, leading to inconsistent literature screening. How can I improve its quality?
A: A high-quality, pre-defined protocol is crucial for reproducibility and minimizing bias. Follow these steps and use established checklists to ensure rigor [42] [43].
The diagram below outlines this structured development workflow:
Q: My reaction failed, yielding no product. I suspect a solvent-related issue. How can I systematically diagnose and resolve this?
A: Solvent selection critically impacts reaction success by influencing solubility, reactivity, and even catalyst stability [45] [46]. Errors in measuring reactants or using the wrong reagents are also common causes of failure [16].
The table below summarizes diagnostic checks for solvent-related failures:
Table 1: Diagnostic Checks for Solvent and Reaction Failure
| Checkpoint | Description | Tool/Metric |
|---|---|---|
| Reagent Identity | Confirm the correct chemical was used; e.g., acetic anhydride vs. acetic acid [16]. | Lab inventory log, CAS numbers. |
| Partition Coefficient (K) | Determine the ideal K value for efficient liquid-liquid extraction during workup [45]. | "Shake-flask" method with HPLC/UV-vis analysis. |
| Solvent Hydrophobicity (log P) | Predict solvent compatibility with enzymes or proteins to prevent denaturation [46]. | Calculated or experimentally determined log P values. |
| Phase Confusion | Ensure the correct aqueous or organic phase is collected during workup [16]. | Knowledge of solvent densities. |
Q: Are my chemical testing methods still compliant with current international regulatory standards?
A: Regulatory guidelines are continuously updated. For 2025, the OECD has introduced significant revisions to its Chemical Testing Guidelines [47].
Table 2: Selected OECD Test Guideline Updates (2025)
| Test Number | Focus Area | Key Update in 2025 |
|---|---|---|
| Test No. 254 | Environmental Toxicity | New guideline for acute contact toxicity testing on solitary bees (Osmia sp.) [47]. |
| Test No. 497 | Skin Sensitization | Now includes Defined Approaches (DAs) integrating in chemico and in vitro data [47]. |
| Test No. 203, 210, 236 | Fish Toxicity | Revised to allow tissue sampling for omics analysis for a deeper understanding of biological responses [47]. |
| Test No. 407, 408, 421, 422 | Repeated Dose & Reproductive Toxicity | Revised to allow tissue sampling for omics analysis [47]. |
Q: What are the critical 2025 regulatory changes for Point-of-Care Testing that my diagnostic lab must implement?
A: CLIA (Clinical Laboratory Improvement Amendments) regulations for POCT were updated in 2025, with a sharper focus on accuracy and personnel qualifications [48].
Table 3: Essential Reagents and Materials for Diagnostic Protocols
| Item | Function/Application | Key Consideration |
|---|---|---|
| HEMWat Solvent System | A versatile, adjustable family of solvent systems for countercurrent separation (CCS) of natural products like terpenoids and flavonoids [45]. | Covers a wide polarity range; proportions can be tuned to achieve ideal partition coefficients (K) [45]. |
| Solvents with log P > 4 | Organic solvents for biocatalysis or protein extraction (e.g., 1-octanol, hexane). | High log P minimizes protein denaturation by reducing solvent interaction with water and the protein's hydrophobic domains [46]. |
| Immunoglobulin G (IgG) | A model protein for developing and validating bioseparation protocols, such as carrier-mediated extraction [46]. | Stability varies significantly with different solvents and ionic liquids; a benchmark for method robustness [46]. |
| PRISMA-P Checklist | A structured tool for writing high-quality systematic review protocols [42]. | Ensures all essential elements of a rigorous diagnostic protocol are included, reducing bias [42]. |
| OECD Test Guidelines | The international gold standard for chemical safety assessment methods [47]. | Required for regulatory compliance and acceptance of data across member countries; must be kept up-to-date with annual revisions [47]. |
FAQ: What is the primary purpose of evaluating a reaction's substrate scope? The primary purpose is to define the boundaries and utility of a synthetic method. A well-designed substrate scope identifies which molecular features (steric, electronic) a reaction is sensitive to, reveals its functional group tolerance, and provides insight into the reaction mechanism. This goes beyond a simple demonstration of utility and enables researchers to predict how new, untested substrates might perform [49].
FAQ: My reaction works for my model substrate but fails for most analogs. What should I troubleshoot? This often indicates that the model substrate was not representative of the broader experimental space. First, analyze the structural differences between your working and failed substrates. Use quantitative molecular descriptors (like Sterimol parameters and IR stretching frequencies) to map the steric and electronic landscape you are trying to access. Your initial substrate scope may be too narrow; you likely need to go back and design a more systematically varied, smaller library to identify the specific structural feature causing the failure [49].
FAQ: How can I design a substrate scope that is informative but not excessively large? Employ principles of Design of Experiments (DoE). Rather than testing many similar substrates, select a smaller set that broadly and systematically samples the key steric and electronic variations you hypothesize will affect the reaction. This involves:
FAQ: What are the common pitfalls in interpreting co-occurrence patterns in complex systems like soil microbiology, and how do they relate to substrate scope evaluation? A major pitfit is drawing causal conclusions from correlative data. In soil microbial networks, the observed co-occurrence of organisms does not necessarily prove a direct biological interaction, as it can be influenced by environmental factors. Similarly, in substrate scope evaluation, the success or failure of a series of substrates might be incorrectly attributed to a single obvious structural feature, when in reality, it is the result of a complex interplay of multiple steric and electronic factors. The conclusions must be limited to the scientific evidence the experimental design can actually provide [50].
| Problem | Possible Cause | Solution |
|---|---|---|
| Low Functional Group Tolerance | The reaction conditions (e.g., catalyst, solvent, pH) are incompatible with certain functional groups, leading to side reactions. | Systematically test the limiting functional group in isolation. Modify reaction conditions to be milder, or use protecting group strategies. |
| Inconsistent Results with Sterically Similar Substrates | Underappreciated electronic differences between substrates are affecting the reaction. | Expand your analysis to include quantitative electronic descriptors, such as the carbonyl IR stretching frequency, to better differentiate between substrates [49]. |
| Inability to Predict Performance of New Substrates | The initial substrate scope was qualitative and lacked systematic variation, making quantitative prediction impossible. | Redesign the substrate library using a DoE approach. Develop a linear regression model that correlates molecular descriptors to reaction outcomes to enable prediction for new substrates [49]. |
| Poor Conversion for Sterically Hindered Substrates | The active site of a catalyst or enzyme cannot accommodate large substituents. | If using an enzyme, explore directed evolution or site-directed mutagenesis to engineer a more open active site, as demonstrated with thermostable phenolic acid decarboxylases [51]. For synthetic catalysts, consider a catalyst with a larger ligand pocket. |
| Trade-off in Activity when Expanding Scope | Optimizing the system for one type of substrate (e.g., sterically hindered) reduces its efficiency for another (e.g., electron-poor). | This is a common challenge. Use the quantitative model to understand the trade-off and identify a balanced set of conditions, or develop two specialized sets of conditions for different substrate classes [51] [49]. |
This methodology enables the development of a predictive, mathematically grounded substrate scope [49].
Identify Parameters and Define Experimental Space:
Organize and Evaluate a DoE-Based Library:
Connect Descriptors to Outcomes via Linear Regression:
ÎÎGâ¡ = -RTln(er)) that links the molecular descriptors (Sterimol parameters, IR frequency) to the experimental reaction outcomes.Apply the Model to Predict New Substrates:
This protocol is adapted from work on thermostable phenolic acid decarboxylases [51].
Library Generation:
High-Throughput Screening:
Characterization of Hits:
Mechanistic Investigation:
The following table summarizes key quantitative descriptors for a representative set of ketone substituents, as used in a DoE-based substrate scope study [49].
Table 1: Sterimol Parameters and IR Stretching Frequencies for Ketone Substituent Selection
| Substituent Category | Example R Groups | Sterimol B1 (à ) | Sterimol B5 (à ) | Calculated Carbonyl IR Stretch (cmâ»Â¹) |
|---|---|---|---|---|
| Small Alkyl | Methyl, Ethyl | 1.52 - 2.90 | 3.00 - 3.82 | 1785 - 1795 |
| Branched Alkyl | i-Propyl, Cyclopropyl | 2.04 - 2.90 | 4.08 - 4.68 | 1786 - 1794 |
| Oxygenated Alkyl | CH2OMe, CH2OEt | 2.67 - 2.90 | 4.89 - 5.50 | 1787 - 1794 |
| Halogenated Alkyl | CH2F, CH2Cl | 2.67 - 2.90 | 4.26 - 4.89 | 1792 - 1800 |
| Small Aryl | Phenyl, 2-Thienyl | 2.90 - 3.20 | 5.50 - 6.20 | 1790 - 1800 |
| Substituted Aryl | 4-MeO-C6H4, 4-CF3-C6H4 | 2.90 - 3.20 | 5.50 - 6.20 | 1788 - 1805 |
Table 2: Key Research Reagent Solutions for Substrate Scope and Mechanism Analysis
| Reagent / Material | Function in Experiment |
|---|---|
| Sterimol Parameters | Quantitative, multidimensional descriptors of substituent steric bulk (B1 for minimum width, B5 for maximum width) used to correlate structure with reactivity [49]. |
| Carbonyl IR Stretch Frequency | A quantitative descriptor of a ketone's electrophilicity, useful for predicting reactivity and modeling transition states [49]. |
| Combinatorial Active Site Library | A collection of enzyme mutants created via mutagenesis of specific active site residues, used to rapidly screen for altered substrate scope or improved activity [51]. |
| Linear Regression Modeling | A statistical method used to build a quantitative relationship between molecular descriptors (e.g., Sterimol, IR frequency) and experimental reaction outcomes (e.g., enantioselectivity) [49]. |
| Molecular Dynamics Simulations | Computational simulations used to visualize and understand how mutations in an enzyme affect active site dynamics, substrate binding, and ultimately, function [51]. |
Diagram 1: Troubleshooting workflow for failed substrate scope.
Diagram 2: Enzyme engineering to expand substrate scope.
Q1: What is reaction optimization and why is it a critical skill for synthetic chemists? Reaction optimization is the systematic process of adjusting experimental conditions to improve the outcome of a chemical reaction [52]. Key outcomes targeted for improvement include chemical yield, conversion, selectivity, and reaction rate [52]. This process is fundamental because published procedures do not always work in a different lab context, and even small changes to conditions can dramatically improve performance, saving significant time, reagents, and money [52]. It is a critical step before scaling up a synthesis for production [52].
Q2: What are the primary variables a researcher can manipulate during optimization? The most common variables are categorized in the table below [52]:
| Variable | Example Options |
|---|---|
| Solvent | MeCN, THF, EtOH, DMSO |
| Temperature | Room temp, reflux, cryogenic (-78 °C) |
| Catalyst | Pd(PPhâ)â, NiClâ(dppp), CuI |
| Reaction Time | 30 min, 4 h, overnight |
| Stoichiometry | Equivalents of reagents, limiting agent |
| Additives (Base/Acid) | KâCOâ, DBU, HCl |
| Concentration | Dilution level (e.g., 0.1 M vs. 1.0 M) |
Q3: What are the main methodological approaches to optimization? The current state of optimization includes several methodologies, ranging from traditional to modern [53]:
Q4: How can I start optimizing a reaction without expensive automated equipment? A step-by-step approach for beginners is both practical and effective [52]:
This guide addresses common experimental observations, their possible causes, and proposed solutions to help diagnose and fix failed reactions.
| Observation | Possible Cause | Solution |
|---|---|---|
| No Product Formed | Incorrect annealing temperature (for PCR) | Recalculate primer Tm values; test an annealing temperature gradient [55]. |
| Poor primer design | Verify primers are non-complementary; increase primer length [55]. | |
| Suboptimal reaction conditions | Optimize Mg²⺠concentration; thoroughly mix all components [55]. | |
| Poor template quality | Analyze DNA via gel electrophoresis; check 260/280 ratio [55]. | |
| Insufficient number of cycles | Rerun the reaction with an increased number of cycles [55]. | |
| Multiple or Non-Specific Products | Primer annealing temperature too low | Increase the annealing temperature [55]. |
| Premature replication | Use a hot-start polymerase; set up reactions on ice [55]. | |
| Poor primer design | Check primer design; avoid GC-rich 3' ends [55]. | |
| Incorrect Mg²⺠concentration | Adjust Mg²⺠concentration in 0.2-1 mM increments [55]. | |
| Incorrect Product Size | Mispriming | Verify primers have no additional complementary regions within the template DNA [55]. |
| Incorrect annealing temperature | Recalculate primer Tm values using a trusted calculator [55]. | |
| Sequence Errors | Low fidelity polymerase | Choose a higher fidelity polymerase [55]. |
| Suboptimal reaction conditions | Reduce number of cycles; decrease extension time; decrease Mg²⺠concentration [55]. | |
| Unbalanced nucleotide concentrations | Prepare fresh deoxynucleotide mixes [55]. |
| Item | Function |
|---|---|
| High-Fidelity DNA Polymerase (e.g., Q5) | Reduces sequence errors in amplified products by providing superior accuracy [55]. |
| Hot-Start Polymerase | Prevents premature replication (primer-dimer formation) by requiring thermal activation, improving specificity and yield [55]. |
| PreCR Repair Mix | Repairs damaged template DNA before amplification, which can help recover product from suboptimal templates [55]. |
| GC Enhancer | A specialized additive that improves the amplification of GC-rich templates, which are often difficult to replicate [55]. |
| Monarch Spin PCR & DNA Cleanup Kit | Purifies the reaction product or starting template to remove inhibitors like salts or proteins that can cause reaction failure [55]. |
This protocol provides a generalized methodology for beginning the optimization process for a new or underperforming organic reaction [52].
Objective: To systematically identify the most impactful variables and establish a baseline of improved performance.
Step 1: Define the Optimization Goal and Metric
Step 2: Literature Review and Hypothesis Generation
Step 3: Design the Initial Experiment Set
Step 4: Execution and Data Collection
Step 5: Data Analysis and Iteration
The following diagram illustrates the logical workflow and iterative nature of the reaction optimization process.
Machine learning (ML) represents the frontier of data-driven condition prediction. A seminal 2018 study demonstrated a neural-network model trained on approximately 10 million reactions from Reaxys to predict suitable chemical context (catalyst, solvent, reagent) and temperature [54]. The model's performance highlights both the potential and current state of ML in the field [54] [53]:
The following diagram outlines the high-level architecture of a neural network for predicting reaction conditions, mapping the input of a reaction to the output of suggested parameters.
Q1: What is AIQM2 and how does it improve upon traditional computational methods? AIQM2 is the second generation of the AI-enhanced quantum mechanics method. It utilizes a delta-learning framework to correct a modified GFN2-xTB baseline to the high-fidelity CCSD(T)/CBS level of theory. This allows it to bypass the accuracy limitations of common Density Functional Theory (DFT) approaches with double-zeta quality basis sets, while operating at the computational cost of semi-empirical methods. It is particularly noted for its superior performance in describing transition states and barrier heights, often achieving chemical accuracy for organic molecules containing CHNO elements [56] [57].
Q2: My geometry optimization for a transition state is failing. What could be wrong? Failed transition state optimizations are often due to an inaccurate initial guess structure. AIQM2 requires a reasonable starting geometry to converge successfully.
ts keyword in your input file to specify a transition state search [57].
Q3: Why does my reaction simulation yield unexpected products, and how can AIQM2 help? Unexpected products in reaction simulations can stem from inaccurate potential energy surfaces that fail to capture the correct reaction pathways, such as bifurcating reactions. Traditional DFT methods at the B3LYP-D3/6-31G* level can be slow and less accurate, leading to incorrect product distribution predictions. AIQM2 addresses this with superior speed and accuracy, enabling the propagation of thousands of trajectories to reliably revise and predict product distributions [56].
Q4: How do I set up a reactive molecular dynamics simulation with AIQM2? Reactive molecular dynamics allows you to simulate chemical reactions in real-time.
Q5: Can I simulate spectra using AIQM2? Yes, AIQM2 can compute molecular properties for spectroscopic simulations, such as infrared (IR) spectra.
Q6: My calculation failed due to charge/multiplicity errors. How do I correct this? This error occurs when the system's charge and spin multiplicity are not defined correctly.
charge and multiplicity properties for your molecule. For input files, use the charges and multiplicities keywords. In a Python script, define these properties directly in the molecule object [57]:
The table below summarizes specific issues and their resolutions.
| Error / Issue | Possible Cause | Solution |
|---|---|---|
| Installation Failure | Missing DFT-D4 dependency. | Install DFT-D4 from PyPI and set the dftd4bin environment variable to the binary path [57]. |
| Convergence Problems | Poor initial geometry or incorrect method parameters. | Provide a better initial structure and ensure the ts=True keyword is used for transition state searches [57]. |
| Incorrect Reaction Path | Low-level of theory (e.g., certain DFT functionals) providing an inaccurate potential energy surface. | Use AIQM2 for a more accurate potential energy surface, which is critical for correct reaction dynamics [56]. |
| Low Accuracy for Non-CHNO Elements | AIQM2 is parameterized for organic molecules containing C, H, N, O. | Use AIQM2@UAIQM for systems containing other elements (excluding 7th-row elements) [57]. |
This is the most basic calculation to obtain the energy of a molecular structure.
Detailed Methodology:
sp.inp):
sp.inp file on the Aitomistic Hub or via the command line with MLatom.This protocol is essential for locating and characterizing transition states, which are critical for understanding reaction kinetics.
Detailed Methodology:
opt.py [57]:
This protocol is used to simulate the real-time dynamics of a chemical reaction.
Detailed Methodology:
This section details the essential software and computational "reagents" required to perform simulations with AIQM2.
| Item Name | Function / Role | Availability / Installation |
|---|---|---|
| MLatom | The open-source computational chemistry package that serves as the primary platform for running AIQM2 calculations. | Available via GitHub: https://github.com/dralgroup/mlatom or PyPI [57]. |
| DFT-D4 | A program for calculating dispersion corrections, which is a required dependency for AIQM2. | Available via PyPI: https://github.com/dftd4/dftd4. The dftd4bin environment variable must be set after installation [57]. |
| Aitomistic Hub | An online platform (XACS cloud) that provides a web-based interface for running AIQM2 simulations without local installation. | Accessible at: https://XACScloud.com [56] [57]. |
| Aitomia | An AI assistant integrated into the Aitomistic Hub that can autonomously derive reaction properties. | Accessed via the "Chat with chatbot" panel on the Aitomistic Hub [57]. |
| Script Keyword / Command | Function in Experiment |
|---|---|
AIQM2 |
Specifies the use of the AIQM2 method for the calculation [57]. |
ts |
A keyword that triggers a transition state geometry optimization instead of a ground state optimization [57]. |
freq |
Requests a frequency calculation following a geometry optimization to characterize stationary points [57]. |
ir |
Keyword used to initiate a calculation for generating infrared (IR) spectra [57]. |
ml.optimize_geometry(ts=True) |
The Python API function and argument for performing a transition state search [57]. |
Q1: My reaction with an informer library failed to provide any diagnostic results. What are the first things I should check? Begin by verifying the integrity of your informer library compounds using the analytical methods outlined in the procurement documentation. Confirm that your reaction conditions are sufficiently diverse to provoke a range of outcomes; if conditions are too specific, they may not engage the diverse functional groups within the library. Finally, ensure your analytical techniques (e.g., LC-MS) are sensitive enough to detect minor products and have been calibrated correctly [58].
Q2: The analysis software does not recognize a fragment from my informer library reaction. How can I proceed? First, manually calculate the molecular formula and mass of the suspected fragment to verify the software's output. Use the mass fragmentation tool in your drawing software to simulate possible fragmentation patterns and compare them with your experimental data. If the fragment remains unidentified, consult the specific informer library's documentation for known decomposition pathways or byproducts associated with its complex scaffolds [58].
Q3: How can I improve the publication-quality of my figures involving complex informer library structures? Utilize the advanced coloring and alignment features in modern chemistry drawing software. Apply ring fill coloring to direct focus to specific parts of a molecule that reacted. Use the alignment tools to produce clean, consistent figures for publication. You can precisely control object colors by entering exact hex codes to ensure visual clarity and meet journal requirements [58].
Q4: I need to modify a complex monomer within a biopolymer for my informer library. What is the most efficient method? Use the find and replace capability within the biopolymer (HELM) editor of your chemical drawing software. This allows you to identify and select specific monomers on the canvas and replace them in bulk, streamlining large-scale modifications of sequences. This is particularly useful for rapidly transforming a natural sequence from a FASTA string into a highly complex sequence with custom modifications [58].
| Scenario | Possible Cause | Recommended Action |
|---|---|---|
| No reaction observed | Informer library compound degradation; inappropriate reaction conditions. | Re-run quality control on library; diversify reaction parameters (catalyst, solvent, temperature). |
| Unidentifiable spectral data | Software misassignment; novel fragmentation pattern; sample impurity. | Perform manual spectral validation; use mass fragmentation simulation tools; re-purify sample. |
| Inconsistent results | Human error in complex setup; reagent decomposition; water/oxygen sensitivity. | Automate liquid handling where possible; use fresh reagents; rigorously exclude air/moisture. |
| Failed informer library synthesis | Incompatible protecting groups; poor functional group tolerance in key step. | Re-evaluate synthetic route using retrosynthetic analysis software; employ orthogonal protection. |
The following table details essential materials and their functions for experiments utilizing chemistry informer libraries.
| Item | Function / Application |
|---|---|
| ChemDraw Desktop & Cloud Applications | Transforms chemical drawings into knowledge, combining advanced drawing capabilities with cloud-native applications for streamlined communication [58]. |
| Mass Fragmentation Tool | Mimics Mass Spec fragmentation to generate fragment structures with calculated molecular formulas and masses, crucial for analyzing reaction outcomes [58]. |
| HELM (Hierarchical Editing Language for Macromolecules) Editor | Specialized tool for editing biopolymers, supporting operations like finding and replacing monomers, which is essential for working with complex biomolecular informer sets [58]. |
| Periodic Table Tool | Allows for the selection of any element to add to the canvas and the creation of atom lists for generating generic structures in informer libraries [58]. |
| Analysis Panel | Displays basic chemical properties (e.g., molecular weight, formula, exact mass) for structures and biopolymers, context-sensitive to the selection on the canvas [58]. |
The table below summarizes hypothetical quantitative properties for a standard set of informer library compounds, illustrating the data points researchers should monitor.
| Compound ID | Molecular Weight (g/mol) | Molecular Formula | Exact Mass | Calculated Log P | Polar Surface Area (à ²) |
|---|---|---|---|---|---|
| INF-CORE-01 | 347.41 | CââHââNOâ | 347.1158 | 3.2 | 75.6 |
| INF-CORE-02 | 285.33 | CââHââ NâOâ | 285.1212 | 2.8 | 89.4 |
| INF-CORE-03 | 432.51 | CââHââNâOâ | 432.1434 | 1.5 | 112.0 |
| INF-CORE-04 | 398.44 | CââHââNâOâ | 398.1267 | 2.1 | 98.2 |
| INF-CORE-05 | 511.58 | CââHââNOâ | 511.1420 | 4.5 | 87.3 |
1. Objective To assess the functional group tolerance and potential side-reactivities of a novel catalytic reaction by challenging it with a diverse set of complex molecular fragments present in a chemistry informer library.
2. Materials and Equipment
3. Procedure Step 3.1: Experimental Setup. In a controlled atmosphere (e.g., nitrogen glovebox), prepare a series of reaction vials. Using an automated dispenser or calibrated pipettes, add a stock solution of each informer library compound (typically 1.0 µmol per compound) to its respective vial.
Step 3.2: Reaction Initiation. Add the standard reaction componentsâcatalyst, base, and solventâto each vial according to the general reaction scheme. Initiate the reaction simultaneously for all vials by placing them in a pre-heated aluminum block stirrer.
Step 3.3: Quenching and Sampling. After the designated reaction time (e.g., 4 hours), quench each reaction simultaneously by transferring a small aliquot into a predefined well of a 96-well plate containing a quenching solution. Dilute the aliquots appropriately for analysis.
Step 3.4: LC-MS Analysis. Analyze each quenched sample using a standardized LC-MS method. The method should be optimized to separate starting materials, desired products, and potential byproducts.
Step 3.5: Data Processing.
Integrate the UV (e.g., 254 nm) peak areas for the starting informer compound and the proposed product in each sample. Calculate the conversion percentage for each informer compound as: [1 - (Area_starting_material / Area_total)] * 100%.
4. Data Analysis
Q1: My calculated reaction energies seem physically unrealistic. How can I determine if the error comes from the electronic structure method itself? A key diagnostic is to check the internal consistency of your calculation. For Coupled Cluster calculations, compute the T1 diagnostic and the non-Hermiticity diagnostic of the one-particle reduced density matrix. The extent of asymmetry in the density matrix indicates how far your calculation is from the exact limit; larger values signal potential inaccuracies [59]. For DFT methods, significant errors can arise from using an inaccurate electron density (density-driven errors). Running a Hartree-Fock DFT (HF-DFT) calculation, where the HF density is used instead of the self-consistent DFT density, can help isolate and reduce these errors [13].
Q2: When troubleshooting a failed catalytic cycle simulation, my DFT and Coupled Cluster results disagree significantly. Which should I trust? This discrepancy often indicates a system with strong electron correlation effects, which are challenging for standard single-reference methods. In this case:
Q3: What specific diagnostic can tell me both how difficult my molecular system is AND how well my computational method is handling it? The non-Hermiticity diagnostic derived from the asymmetry of the one-particle reduced density matrix in Coupled Cluster theory provides this dual insight [59]. It is calculated as the Frobenius norm of the anti-symmetric part of the matrix, normalized by the square root of the number of electrons. A larger value indicates a more challenging system (e.g., with multireference character), while a reduction in this value when you use a higher-level CC method (e.g., moving from CCSD to CCSDT) shows that the method is improving the description [59].
Follow this logical workflow to diagnose the source of inaccuracies in your reaction energy calculations.
Troubleshooting Workflow for Reaction Energy Accuracy
| Diagnostic | Method | Formula / Principle | Acceptable Range | Indication of Problem |
|---|---|---|---|---|
| T1 Diagnostic [59] | Coupled Cluster (e.g., CCSD) | Norm of single excitation vector | < 0.05 | High multireference character |
| Non-Hermiticity Diagnostic [59] | Coupled Cluster | ( ||Dp^q - {Dp^q}^T||F / \sqrt{N{\text{electrons}}} ) | Lower is better, 0 is exact | Method inadequacy & problem difficulty |
| Density-Driven Error Check [13] | Density Functional Theory | Compare self-consistent DFT vs. HF-DFT energy error | HF-DFT error significantly smaller | Inaccurate self-consistent density |
| Reagent / Material | Function in Computational Validation |
|---|---|
| Coupled-Cluster Theory | Provides systematically improvable, high-accuracy reference energies for benchmarking [59]. |
| Hartree-Fock Density | Used in HF-DFT to isolate and correct density-driven errors in DFT calculations [13]. |
| One-Particle Reduced Density Matrix | Core quantity for calculating properties and the non-Hermiticity diagnostic in CC theory [59]. |
| Lambda ((\Lambda)) Operator | Key component for left-hand Coupled Cluster wavefunction and density matrix calculation [59]. |
Purpose: To assess the reliability of Coupled Cluster reaction energies and identify potential multireference issues.
Methodology:
Purpose: To determine if inaccuracies in DFT reaction energies stem from flaws in the self-consistent electron density.
Methodology:
This troubleshooting guide focuses on three powerful approaches for optimizing organic reactions and troubleshooting failures:
High-Throughput Experimentation (HTE): Uses automation to rapidly conduct hundreds or thousands of parallel experiments, generating extensive data on reaction parameters and outcomes. This is particularly valuable for exploring complex, multi-variable reaction spaces efficiently.
Design of Experiments (DoE): A statistical approach that systematically varies multiple factors simultaneously to identify optimal conditions and understand factor interactions. This method helps researchers move beyond inefficient one-factor-at-a-time (OFAT) approaches.
Machine Learning (ML): Applies algorithms to analyze complex datasets, predict optimal reaction conditions, and identify patterns that may not be apparent through traditional analysis. ML models can become increasingly accurate as more experimental data is accumulated.
Failed reactions contribute significantly to the high failure rates in drug development. Approximately 40-50% of clinical drug development failures are attributed to lack of efficacy, while 30% result from unmanageable toxicity [60]. Efficient reaction optimization methodologies directly address these challenges by ensuring the reliable synthesis of compounds with desired therapeutic properties. The persistent 90% failure rate in clinical drug development underscores the importance of robust experimental design in early research stages [61] [60].
Q1: Our HTE screening results show poor reproducibility between microtiter plates. What could be causing this?
Q2: How can we handle the large data volumes generated by HTE without becoming overwhelmed?
Q1: Our DoE models have poor predictive power despite statistical significance. What's wrong?
Q2: How should we handle categorical variables (e.g., catalyst types) in DoE?
Q1: Our ML models perform well on training data but poorly on new reactions. Why?
Q2: How much data is needed to build effective ML models for reaction prediction?
| Characteristic | HTE | DoE | ML |
|---|---|---|---|
| Best Application Scope | Broad condition screening with available automation | Understanding factor interactions with limited experiments | Complex pattern recognition in large datasets |
| Experimental Efficiency | High throughput (100-10,000 experiments) | Medium efficiency (20-100 experiments) | Variable (improves with more data) |
| Data Requirements | Large experimental datasets | Structured experimental designs | Large, high-quality datasets |
| Implementation Complexity | High (requires specialized equipment) | Medium (requires statistical expertise) | High (requires data science expertise) |
| Interpretability | Direct experimental observation | Clear factor-effect relationships | Often "black box" without explanation methods |
| Optimal Project Stage | Early exploration of vast condition spaces | Systematic optimization of key parameters | Late-stage optimization with sufficient historical data |
| Resource Requirements | High equipment cost, moderate personnel | Low equipment cost, high statistical expertise | Variable computing resources, high expertise |
Q1: What's the most efficient sequence for applying these methodologies?
Q2: How can we validate that our optimization approach is working effectively?
Q1: How do we ensure our optimization algorithms perform consistently across different reaction classes?
Q2: What tools are available for managing the computational complexity of these approaches?
| Tool Category | Example Tools | Primary Function | Implementation Consideration |
|---|---|---|---|
| DoE Software | JMP, Design-Expert, Modde | Experimental design creation and analysis | Requires statistical expertise for proper implementation |
| HTE Platforms | Chemspeed, Unchained Labs | Automated reaction setup and analysis | High initial investment, specialized maintenance |
| ML Libraries | Scikit-learn, DeepChem, RDKit | Algorithm implementation for chemical data | Python proficiency required, varying computational demands |
| Data Management | Electronic Lab Notebooks (ELNs), CSD, PDB | Structured data storage and retrieval | Critical for reproducible research and model training |
Q1: Why do promising preclinical results often fail to translate to clinical success?
Q2: How can we better prioritize reaction optimization efforts in drug development pipelines?
Q1: Why do reactions optimized at small scale sometimes fail during scale-up?
Q2: How can we build scalability into our initial optimization strategies?
Q1: What is the core concept behind "tera-scale data mining" for organic chemistry research? A1: Tera-scale data mining refers to the use of sophisticated machine learning algorithms to systematically re-analyze vast existing archives of experimental data (often spanning terabytes of stored high-resolution mass spectrometry (HRMS) files) to test new chemical hypotheses and revise reaction mechanisms. This approach, often called "experimentation in the past," repurposes previously acquired but under-analyzed data, potentially revealing novel transformations and insights without the need for new laboratory experiments [62].
Q2: My reaction failed. Beyond checking my starting materials, how can data mining help diagnose the issue? A2: Machine learning-powered search engines like MEDUSA Search can scour your historical HRMS data for specific ion signatures [62]. This allows you to:
Q3: What are the common data quality issues when applying machine learning to existing reaction data? A3: Success depends heavily on data quality and management. Key challenges include [64] [65]:
Q4: Which machine learning models are best suited for analyzing reaction data? A4: The choice of model depends on the specific task:
Problem: Your analysis of DDA or DIA mass spectrometry data is yielding a low number of identified peptides or small molecule compounds.
| Possible Cause | Diagnostic Steps | Solution |
|---|---|---|
| Low Signal-to-Noise Ratio | Inspect raw spectra for high baseline noise. Check peak intensity thresholds in your processing software. | Optimize MS instrument settings. Use a deeper learning-based feature detection algorithm that is more robust to noise [68]. |
| Suboptimal Data Analysis Workflow | Compare the number of identifications from a spectral library search versus a database search. | Implement an integrated analysis workflow that combines spectral library search, database search, and de novo sequencing to maximize sensitivity, as demonstrated by platforms like PEAKS Online [68]. |
| High False Discovery Rate (FDR) | Check the FDR reported by your search engine. Examine the confidence scores of identified spectra. | Employ a unified FDR framework that uses a consistent target-decoy approach across all search methods. Utilize deep learning-based spectrum prediction to rescore and validate identifications [68]. |
Problem: Your automated search of tera-scale MS data archives returns many incorrect hits (false positives) or misses known compounds (false negatives).
Workflow for Diagnosing Search Algorithm Performance:
Solutions:
Problem: Processing thousands of samples and terabytes of data is impractically slow on local servers.
Guidance:
This protocol is adapted from the highly sensitive workflow for immunopeptidomics, which can be applied to other areas requiring deep coverage [68].
Quantitative Performance Comparison of DIA Analysis Tools: Table: Number of Identified Peptide Precursors at 1% FDR on Benchmark DIA Datasets [68]
| Benchmark Dataset | PEAKS Online | DIA-NN | Spectronaut |
|---|---|---|---|
| Muntel et al. | ~105-130% (Baseline) | ~100% (Baseline) | ~100% (Baseline) |
| Xuan et al. | ~105-130% (Baseline) | ~100% (Baseline) | ~100% (Baseline) |
| ABRF Study | Consistent results across runs | - | - |
This protocol, based on the MEDUSA Search engine, enables the discovery of new reactions from existing data [62].
Key metrics to evaluate before proceeding with tera-scale data mining: [62] [64]
Table: Essential Computational Tools for Tera-Scale Data Mining in Organic Chemistry
| Tool / Resource | Function | Application in Research |
|---|---|---|
| MEDUSA Search [62] | Machine Learning-Powered Search Engine | Discovers unknown reactions and intermediates in archived HRMS data by searching for specific isotopic patterns. |
| PEAKS Online [68] | Streamlined MS Data Analysis Platform | Integrates DDA/DIA analysis, database/search/library search, and de novo sequencing via deep learning for highly sensitive peptide/compound identification. |
| Vertex AI Pipelines [65] | ML Workflow Orchestration | Formalizes, automates, and monitors end-to-end ML workflows (e.g., data prep, training, serving) for reproducibility and scalability on cloud infrastructure. |
| Chemma LLM [67] | Fine-Tuned Large Language Model | Assists in retrosynthesis planning, reaction yield prediction, and condition generation by learning from vast reaction databases (e.g., USPTO-50k). |
| Graph Neural Networks (GNNs) [66] | Molecular Representation & Prediction | Represents molecules as graphs for highly accurate prediction of reaction outcomes, properties, and optimization. |
| BRICS Fragmentation [62] | Retrosynthetic Fragmenter | Automatically generates hypothetical molecular fragments for constructing query ions in a reaction discovery search. |
The landscape of organic reaction troubleshooting is undergoing a profound transformation, moving from reliance on chemical intuition alone to integrated approaches that combine foundational knowledge with cutting-edge technologies. The synergy of High-Throughput Experimentation, Machine Learning prediction models, automated optimization platforms, and advanced computational methods creates a powerful toolkit for addressing reaction failures systematically. These approaches enable researchers to navigate complex parameter spaces efficiently, predict suitable conditions for novel transformations, and validate solutions against pharmaceutically relevant complexity. For biomedical and clinical research, adopting these methodologies promises to accelerate synthetic routes to target molecules, expand accessible chemical space for drug candidates, and ultimately reduce the timeline from discovery to development. Future directions will likely focus on increased integration of AI-driven discovery with automated validation, creating fully autonomous systems for reaction development and optimization that further democratize access to robust synthetic methodologies across the research community.