Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Micro- and milli-fluidic sample environments for in situ X-ray analysis in the chemical and materials sciences

Mark A. Levenstein *, Corinne Chevallard , Florent Malloggi , Fabienne Testard and Olivier Taché
Université Paris-Saclay, CEA, CNRS, NIMBE, LIONS, 91191, Gif-sur-Yvette, France. E-mail: mark.levenstein@cea.fr

Received 30th July 2024 , Accepted 4th November 2024

First published on 8th January 2025


Abstract

X-ray-based methods are powerful tools for structural and chemical studies of materials and processes, particularly for performing time-resolved measurements. In this critical review, we highlight progress in the development of X-ray compatible microfluidic and millifluidic platforms that enable high temporal and spatial resolution X-ray analysis across the chemical and materials sciences. With a focus on liquid samples and suspensions, we first present the origins of microfluidic sample environments for X-ray analysis by discussing some alternative liquid sample holder and manipulator technologies. The bulk of the review is then dedicated to micro- and milli-fluidic devices designed for use in the three main areas of X-ray analysis: (1) scattering/diffraction, (2) spectroscopy, and (3) imaging. While most research to date has been performed at synchrotron radiation facilities, the recent progress made using commercial and laboratory-based X-ray instruments is then reviewed here for the first time. This final section presents the exciting possibility of performing in situ and operando X-ray analysis in the ‘home’ laboratory and transforming microfluidic and millifluidic X-ray analysis into a routine method in physical chemistry and materials research.


image file: d4lc00637b-p1.tif

Mark A. Levenstein

Mark A. Levenstein is a research scientist at the French Alternative and Atomic Energy Commission (CEA). He completed his PhD at the University of Leeds (UK) on the development of micro- and milli-fluidic devices for synchrotron and laboratory-based X-ray analysis of crystallization processes. After a postdoctoral position at the University of Illinois Urbana-Champaign (USA) on the design of cementitious materials for ecosystem restoration, Mark joined the Interdisciplinary Laboratory on Supramolecular and Nanometric Organization (LIONS). His current research is focused on developing automated synthesis and characterization platforms for the discovery of sustainable materials with applications in energy, electronics, and construction.

image file: d4lc00637b-p2.tif

Corinne Chevallard

Corinne Chevallard is a research director at CEA (France). She holds a PhD in physics from the University of Nice-Sophia-Antipolis (France). Her early research focused on instabilities in complex systems, whether liquid crystals (PhD) or polymers (postdoc at the Weizmann Institute of Science, Israel). Since joining the CEA, her research has focused on inorganic crystallization, particularly in relation to biomineralization. In order to characterize the transient structures formed prior to crystallization, she is developing microfluidic devices specifically designed for synchrotron-based techniques, which enable in situ structural and chemical characterization in liquid media at the nanoscale.

image file: d4lc00637b-p3.tif

Florent Malloggi

Dr. Florent Malloggi received his PhD from Paris Cité University (formerly Paris VII) in 2006, specializing in the physics of liquids. He now serves as a research director in the Nanoscience and Innovation for Materials, Biomedicine and Energy (NIMBE) department of CEA within Paris-Saclay University. His research interests are focused on instrumental microfluidic development for applications in energy, chemistry and biology.

image file: d4lc00637b-p4.tif

Fabienne Testard

Dr. Fabienne Testard is a research director at CEA, working in the LIONS Laboratory of NIMBE. She has expertise in soft-matter, self-assembling systems and nanomaterials with a strong background in synthesis, characterization and the study of the formation mechanisms of nanomaterials (mainly gold and organic self-assembled particles). She is particularly interested by the behavior of nanomaterials in complex systems and in situ SAXS and SANS approaches (with extensive experience at synchrotrons and neutron sources).

image file: d4lc00637b-p5.tif

Olivier Taché

Olivier Taché is a researcher at CEA/NIMBE/LIONS and an expert on the characterization and metrology of nanomaterials. His work includes the development of advanced SAXS methodologies and software to enhance the precision and reliability of nanoparticle size and concentration measurements. He is actively involved in various collaborative projects aimed at enhancing nanosafety and the traceability of nanomaterial measurements.


1 Introduction

There remain many fundamental questions in the chemical and materials sciences relating to natural phenomena and industrial processes including the nucleation and growth pathways of materials,1,2 the function and deactivation of catalysts,3,4 and the operation and failure modes of batteries.5–7 X-rays seem ideally suited to provide insight into these areas since they facilitate a suite of structural, chemical, and imaging techniques, offer high spatial and temporal resolution, and have the penetrating ability required to make in situ or operando observations within thick three-dimensional (3D) samples. These strengths have only increased in recent years due to the widespread availability of third generation synchrotron light sources,8 the arrival of X-ray free electron lasers (XFELs) and fourth generation synchrotrons,9,10 and large improvements in commercial X-ray instruments.11 However, for many types of samples, one bottleneck is introducing the analyte into the X-ray beam in a way that fully exploits these attributes: for example, rapidly and uniformly mixing liquid reagents to initiate a reaction within the short timescales required to make a single measurement.

Microfluidic devices have been proposed as one solution to this sample preparation and manipulation problem. In an excellent earlier review in this journal, Ghazal et al. covered the benefit of microfluidics for X-ray analysis in the life sciences and soft matter research.12 Their review was primarily focused on the ways microfluidic devices could be used to produce and manipulate large numbers of samples and make better use of experimental time at synchrotron facilities. They also highlighted the ways that microfluidic devices could be utilized to perform time-resolved measurements of processes not accessible by conventional macroscale methods. Here, we seek to build upon their work by covering applications in physical chemistry and materials science and focusing primarily on the role of microfluidic devices as “sample environments”. We define these as specialized devices and reactors that enable in situ or operando measurements of samples and processes in their native states or under non-equilibrium conditions. For the purposes of this review, in situ means that a measurement is performed where the reaction or process takes place without moving the sample into a second holder or vessel, and operando means that a measurement is performed both in situ and while the reaction or process is occurring. By this definition, all operando measurements are in situ, but not all in situ measurements are operando. This is not necessarily the same definition as used in catalysis literature.13 The opposite would then be ex situ or post-mortem analysis, where a sample must be moved out of its native environment and prepared for measurement, often requiring quenching, washing, and drying steps, among others.14

A growing interest across the chemical, physical, and natural sciences is the study of processes in real-time. For processes that are difficult to observe in nature (e.g., high-pressure phenomena in the Earth's interior) or in industrial environments (e.g., within large chemical reactors), these conditions need to be re-created in the laboratory. Microfluidic and other miniaturized devices present a promising way to achieve these goals owing to their ability to control heat and mass transport phenomena quickly and precisely and due to their increased safety over macroscale methods when working under extreme temperatures, pressures, and chemical conditions. Here, our analysis is not restricted to microfluidic devices with channel dimensions ≪1 mm, because as many have correctly highlighted,12,15 a longer X-ray beam pathlength through samples results in an optimized signal-to-noise ratio (≈1 mm with aqueous solutions and Cu Kα radiation). Therefore, for some applications, millifluidic sample environments may even be preferred over their microfluidic counterparts.

Despite a handful of earlier papers, such as on the microfluidic preparation of crystals for subsequent off-chip X-ray diffraction,16,17 it may have been Greaves and Manz (2005) who first recognized the potential of microfluidic devices for on-chip chemical X-ray analysis.18 They highlighted, in particular, the power of X-ray fluorescence and X-ray diffraction for elemental analysis and particle identification in flow, respectively, and anticipated that performing real-time measurements on-chip could reduce the time required to optimize crystallization and synthesis conditions. While they did investigate on-chip X-ray generation to make a true “lab-on-a-chip”, they also realized that microfluidic devices could serve as powerful complements to full-scale laboratory equipment, and this is precisely the direction in which the field of microfluidic X-ray analysis has gone over the past 20 years. In fact, in this context, microfluidic devices have been coupled to some of the largest scientific instruments in existence: synchrotron particle accelerators and free-electron lasers. This striking combination of the very big and the very small, already inherent in such facilities, offers the possibility of combining the precise manipulation of subatomic particles with that of molecules and nano- and micro-objects in flow. In addition to this, the control of flow can also solve practical problems like mitigating radiation-induced sample damage and heating, which is becoming more and more important as X-ray sources grow stronger.19

Our goal with this review is to provide as comprehensive of an account as possible into the use of micro- and milli-fluidic sample environments for in situ X-ray analysis in the physical and chemical sciences. Therefore, we will cover applications across the three main types of X-ray techniques: scattering/diffraction (section 3), spectroscopy (section 4), and imaging/tomography (section 5). There are many commonalities in the design considerations and technical challenges of sample environments for these different techniques, and we believe that the communities around each technique could benefit from the sharing of knowledge. Preceding this central part of the review, we will also provide some background on predecessor and alternative sample environments for in situ X-ray studies in liquids and a brief discussion of some important parameters to consider in relation to microfluidic sample environments (section 2).

As already mentioned, due to the need for fast measurements of dynamic and often dilute samples, microfluidic X-ray experiments are typically performed at synchrotron light sources and sometimes XFELs. These facilities are operated on a competitive, proposal-based, user access model where individual beamlines or end-stations dedicated to scattering, spectroscopy, imaging, or a combination of techniques must be solicited for experimental time. Further, these facilities are seldom located near a researcher's home institution. Therefore, experimental “beamtime” is not guaranteed, not limitless, and often not easy in terms of the transport and set-up of complex equipment. Near the end of the review, we will thus cover progress in utilizing micro- and milli-fluidic sample environments with laboratory X-ray instruments (section 6), where techniques such as X-ray diffraction (XRD), X-ray fluorescence (XRF), and micro-computed tomography (μCT) are already readily available at most universities and research institutes. The increasing feasibility of laboratory-based analysis should facilitate much easier and more practical in situ X-ray experiments. It should also allow many more researchers to benefit from fluidic X-ray sample environments in their own research, when the high brilliance or coherence provided by large-scale X-ray facilities is not strictly required. Finally, we will present our perspective on the current state of the field and some practical tips regarding device fabrication and best practices (section 7). This will conclude with a look towards the future focused on promising trends and developments that we think will guide the field over the coming years.

This review will be limited to micro- and milli-fluidic X-ray sample environments where measurements are performed in situ, on-chip, and on inorganic and/or hard condensed matter samples. Articles pertaining to only biological or soft condensed matter samples, such as those already reviewed by Ghazal et al.,12 will be referenced only where they have made an important technical contribution later implemented for physical chemistry or materials research. An exception will be made in section 6 on laboratory-based analysis, where studies of all sample types will be covered, since these have not been reviewed previously. Similarly, articles where analysis is performed on a droplet, jet, or spray exiting a microfluidic device will also be excluded, although some will be introduced briefly in the background section. Such jet-based devices have been covered previously in the context of serial crystallography, and the reader is directed to these reviews for further information.20–22 To the best of our ability, this review is comprehensive up through the year 2023 unless otherwise noted.

2 Background

This background section is not intended to be exhaustive, rather its goal is to provide some general context for the review. It will focus on two important pre-microfluidic technologies and two important microfluidics-parallel technologies, all of which remain in use today and have weaknesses—and also some strengths—compared to microfluidics. This will provide the reader with the setting in which micro- and milli-fluidic sample environments operate and a general knowledge of the types of liquid sample manipulation tools currently available at large-scale X-ray facilities. We will finally present some design considerations and definitions for describing micro- and milli-fluidic flow reactors and their utility for in situ X-ray analysis, which will aid in evaluating and comparing the devices presented in the following review.

2.1 Precursors to microfluidic devices

2.1.1 Stopped-flow devices. Perhaps the most common device for performing operando studies of chemical and biochemical reactions and precipitation processes is the stopped-flow device. It is popular not only for X-ray analysis, but also for neutron scattering,23 FTIR spectroscopy,24 and a range of UV-vis based spectrophotometric and fluorimetric methods,25 to name a few. In a basic stopped-flow configuration, two reactant solutions contained in separate syringes are quickly combined in a turbulent mixing unit and introduced into a thin capillary for observation as a function of reaction time (Fig. 1a).26 The observation should coincide with the rapid stoppage of the flow by, for example, the closure of an electromechanical valve (“hard-stop”) to prevent back-flow into the capillary and to ensure reproducibility in the analysis of reaction kinetics. This method has been widely adopted at synchrotron sources for different techniques (e.g., SAXS, XRD, XAS) to study processes including micellar transformations,27 catalysis,28 the nucleation and growth of nanoparticles,29,30 the early stages of mineral formation,31–33 and the crystallization of metal–organic frameworks (MOFs)34,35 and biochemicals.36
image file: d4lc00637b-f1.tif
Fig. 1 Predecessors and alternatives to micro- and milli-fluidic devices for X-ray analysis. (a) Illustration of a stopped-flow device mounted at an X-ray scattering facility. Two reactants are injected through a mixing element into a capillary where the flow is stopped by a fast valve (adapted with permission from Virtanen et al., 2019; Copyright 2019 American Chemical Society).26 (b) Design of a capillary gas cell with flow and heating capabilities. The inset shows a detailed view of the capillary where the powder sample is placed (adapted with permission from Chupas et al., 2008; International Union of Crystallography).37 (c) A droplet injector for use with XFELs. Each droplet is hit by a single femtosecond X-ray laser pulse (adapted from Roessler et al., 2016; with permission from Elsevier).38 (d) Design of a droplet levitator comprising acoustic transducer arrays to position a liquid sample at a focal point within an X-ray beam (adapted with permission from Morris et al., 2019; CC BY).39

The popularity and longevity of the stopped-flow method likely stem from its accessible operation yet powerful performance. Originally developed in the 1940s, stopped-flow devices required a much smaller volume of solution than their large continuous flow predecessors.25,40 Utilizing passive turbulent mixers, such as the Ball-Berger design,41 stopped-flow devices can also achieve mixing times in the 1 ms to 0.01 ms range (with dead times before observation from ∼10 ms down to 0.1 ms).42 From a practical standpoint, many off-the-shelf commercial devices exist, which are often available at synchrotron beamlines and already integrated with beamline hardware and acquisition software. While they are not always simple to use, they may be easier to work with than many home-made devices.

Despite these numerous strengths, stopped-flow devices have some weaknesses that are especially pronounced in the case of X-ray analysis. One is that the time-resolution of the measurements is limited by the duration of X-ray exposure required to achieve a good signal-to-noise ratio. Taking X-ray scattering as an example, the small number of photons elastically scattered compared to the number of photons in the incident beam often requires the use of high-brilliance synchrotron radiation. However, even at many current beamlines, it is difficult to obtain a good quality small angle X-ray scattering (SAXS) pattern from exposures much shorter than 0.02–1 s (i.e., frame rates of 1–50 Hz) depending on the sample contrast. This is especially true at early reaction times of less than a few seconds, when weakly scattering and/or dilute reaction intermediates—requiring even longer exposures to be detected—are present. Moreover, longer exposure times are also required for performing wide-angle X-ray scattering (WAXS).

One solution to this first weakness of stopped-flow devices is to average the results of several experiments. However, combining individual time-resolved frames from different experiments requires very high reproducibility in the mixing conditions, the cleaning procedure, and the timing between the data acquisition workflow and the operation of the device. This is not always the case due to, for example, the formation of bubbles during mixing and inconsistencies in the mixing ratio.43 Another solution, such as increasing the photon flux, only highlights another weakness of stopped-flow devices. This is that the sample is retained in the capillary and continually exposed to high energy ionizing radiation. Depending on the sensitivity of the sample, it is possible that radiation-induced heating or beam damage could alter the process under study and introduce significant uncertainty into data interpretation. Indeed, this problem can even affect inorganic materials,44,45 and it has only increased at fourth generation synchrotrons.19

2.1.2 Capillary gas (Norby) cells. The second precursor sample environment that we will discuss is the capillary gas cell, also known as the Norby cell. Named after its inventor P. Norby, the original Norby cell is a simple device comprising a glass capillary and a modified T-piece tube fitting (e.g., Swagelok) designed for studying hydrothermal and gas–solid reactions by XRD.46 One end of the metal T-piece is connected to the capillary, and the other two ends are reserved for a connection to a pressurized gas line and for mounting onto a standard goniometer head, respectively. In this way, a sample inside the capillary (typically a powder) can be placed under pressure and easily manipulated within a diffractometer. The capillary can also be externally heated using a hot air blower46 or cooled using a cryogenic gas flow.47 More complex versions of the device that permit continuous gas flows over the powder and that integrate heating elements are now routinely used at X-ray powder diffraction, X-ray absorption spectroscopy, and high energy synchrotron beamlines (Fig. 1b),37 where they are also sometimes known as Clausen cells.48

Capillary gas cells are powerful sample environments that enable a range of in situ and operando studies under extreme conditions. These include investigations of hydrothermal synthesis,49,50 catalysis,48,51 gas capture and storage,52 solid phase transformations,53 and other gas–solid interactions.54 As already stated, their simplicity, ubiquity, and utilization of commercial components make them attractive to many researchers, however, they do have some weaknesses. They are primarily designed to interact with a pre-loaded powder, slurry, or sample bed that is fixed in place. Outside of a few exceptions, in situ generation of the sample (e.g., synthesis from solution) or subsequent manipulation or interaction (e.g., fluid injection) with the sample is not supported. There are versions of the capillary cell that enable high pressure liquid flows, but these are essentially large millifluidic systems55,56—although they will not be treated further here. Even considering these exceptions, reaction products are still not recycled under the beam, presenting the same potential for radiation damage as stopped-flow devices. Additionally, the high temperature and pressure of the gas cells and the fragility of the glass capillaries present a safety risk that must be considered during the experiment.52,55

2.2 Parallel technologies

2.2.1 Liquid sample injectors. The need for high-throughput sample introduction for macromolecular serial crystallography (SX), X-ray photon correlation spectroscopy (XPCS) and other coherent scattering, spectroscopy, and imaging techniques at XFELs has led to the development of a range of liquid sample injectors over the past two decades (Fig. 1c). Although these devices are often considered “microfluidic”, for the purpose of this review, they will be treated separately since analysis is performed off-chip in a free jet or droplet. The high fluence of an XFEL pulse would damage the window of a microfluidic device, rendering it inoperable or degrading the signal quality. Conversely, there are indeed microfluidic devices designed for “fixed-target” serial synchrotron crystallography (SSX), where analysis is performed on-chip. However, these are generally restricted for use in structural biology and are, therefore, out of the scope of this review. Interested readers can find information about devices for fixed-target serial crystallography from other sources.12,21

Injector-based serial femtosecond crystallography (SFX) was first demonstrated by Chapman et al.57 using a gas dynamic virtual nozzle (GDVN),58 which can be tuned to produce jets or monodisperse droplets with the use of a piezoelectric actuator. A very fast (∼10–100 m s−1) and thin (∼1–25 μm diameter) fluid jet is created by a high flow rate of sample (typically protein crystals in their mother liquor) surrounded by a sheath flow of a low density gas (e.g., He) in order to rapidly replenish the sample stream after each X-ray pulse (up to MHz frequency).59 However, for rare or expensive samples the amount of liquid consumed to maintain this jet is too high, and utilizing crystals grown and injected in viscous lipidic cubic phases (LCPs) was later shown to enable jet formation at lower flow rates, reducing sample consumption by a factor of 20.60 Other groups have developed drop-on-demand systems, such as acoustic injectors that can dose nanoliter to picoliter droplets directly from a microwell plate (Fig. 1c).38 While many of the uses of these XFEL injector systems have focused on structural biology, they have also facilitated fundamental physics and chemistry experiments including investigating the ionization61 and supercooling of water,62 ice nucleation,63 the structure of semiconducting microcrystals,64 and the diffusion dynamics of nanoparticles.65

In addition to studying static pre-grown crystals and pure liquids, experimental methods have been developed for operando studies of dynamic reactions and processes. The first is the well-known “pump-probe” method, in which a sample is hit mid-flight with an optical laser pulse and subsequently probed by an X-ray laser pulse after a carefully timed delay.66,67 Of greater interest here, is the so-called “mix-and-inject” method, in which liquid reactants are rapidly mixed and introduced into the XFEL beam.68 The delay time between the mixing point and the point of interaction with the beam determines the time point of the reaction that is probed.69 This technique has primarily been performed at XFEL sources to study the conformational changes of proteins and nucleic acids upon ligand binding.70–72 However, it has also been demonstrated at synchrotron sources,73 where storage ring upgrades74 and the possibility of using high-flux, polychromatic, ‘pink’ beams75 make these experiments more and more feasible.

Injector-based serial crystallography, which has been performed largely at XFELs, and microfluidic X-ray scattering (discussed in section 3), which has been performed primarily at synchrotrons, developed quite independently. However, in recent years there has been more overlap between the two communities as serial crystallography has been increasingly performed at synchrotrons – sometimes even at the same beamlines as on-chip microfluidics experiments.76,77 This seems natural considering the two fields use much of the same equipment, face many of the same technical challenges, and have similar goals, albeit often studying different sample types. A strength of both techniques is that samples are constantly replenished in the beam to minimize the effects of radiation damage on data collection. This is especially true of SFX at XFELs, where femtosecond data collection physically outruns degradation processes (so-called ‘diffraction before destruction’).57 An additional advantage of injector-based techniques compared to on-chip microfluidic analysis is that no device materials are in the beam path to attenuate the signal or produce background noise. Yet, while both types of experiments are complex, currently injector-based XFEL experiments are difficult and require a team of scientists and engineers to perform. Likewise, the requirements for device fabrication and interfacing with XFEL hardware are much stricter than with synchrotron-based microfluidics experiments. Finally, owing to the experimental design and geometry, only a single time-point can be collected per injector-based XFEL experiment. This requires several separate experiments to probe different intermediate states in a reaction, which takes a great deal of time and requires high reproducibility.

2.2.2 Droplet levitators. A less common, but still very useful liquid sample environment is the acoustic droplet levitator (Fig. 1d). Using either a piezoelectric sonotrode (i.e., a Langevin horn)78 or arrays of ultrasonic transducers,39 these devices create standing acoustic waves containing pressure nodes where a liquid droplet of ∼5 nL to 5 μL (∼0.2 to 2 mm diameter) can be levitated.78 In contrast to most other sample introduction techniques, the free-floating droplet is not in contact with any liquid or solid surface, making droplet levitation a powerful technique for the study of bulk phenomena away from surfaces, such as homogenous nucleation.78 This also means that there are no reactor walls or windows for the X-ray beam to pass through, reducing the background noise. The acoustic trapping also imparts some motion to the droplet, which helps to rotate crystals for single-crystal X-ray diffraction (SCXRD) experiments39 and to circulate fluid to avoid radiation damage.79 Compared to most flow-based devices like droplet injectors—and even microfluidic devices—total sample consumption is very low, requiring just a few drops to be deposited with a microliter syringe.80

Despite the many strengths of droplet levitators, there are also some drawbacks, particularly related to performing in situ experiments. After the initial deposition of the sample, controlled mixing and/or subsequent operations on the droplet(s) are difficult. Sample evaporation is also a major problem, unless evaporation is used to initiate the process under study78 or to map a parameter space.79 Evaporation of the solvent concentrates reactants in the droplets, introducing an additional uncontrolled variable into in situ chemistry experiments. There are ways to avoid or minimize sample evaporation, but these each have tradeoffs that compromise other advantages of droplet levitators. For example, aqueous droplets can be covered in an immiscible oil layer that inhibits water transport,39 however this introduces a liquid–liquid interface. Alternatively, large droplets in which evaporative losses are negligible to the total volume can be utilized, but these will be difficult to uniformly mix and may have large inhomogeneities in composition and temperature. Finally, droplets can be levitated in an environment with controlled temperature and humidity to prevent evaporation,80,81 but this normally requires a sample chamber with walls through which the X-ray beam must pass. Therefore, in many cases, droplet levitation may be better for the introduction and manipulation of static samples rather than as an operando X-ray sample environment.

2.3 Micro- and milli-fluidic X-ray sample environments

2.3.1 Advantages. Microfluidic devices and their millifluidic counterparts offer varying advantages compared to the other liquid manipulation approaches discussed above. The general strengths of microfluidic devices have been well documented;82–84 these include low sample volume utilization, minimal risk and severity of sample leakage, precise control of mixing and diffusion, and a high surface area-to-volume ratio that leads to efficient heat transfer, faster reaction kinetics, and other beneficial properties. Micro- and milli-fluidic devices also have several specific advantages as sample environments for in situ X-ray analysis, especially continuous flow devices. The first is simply their small footprint, which makes them easier to mount in a beamline experimental hutch or to fit within a laboratory X-ray instrument that has considerable space restrictions. The second advantage is the rapid replenishment of the sample in the X-ray beam by the flow, which minimizes radiation damage and heating.85 This is a particular advantage micro- and milli-fluidic devices have over stopped-flow devices, Norby cells, and droplet levitators, in which a static sample is held within the beam (N.B., there are also fluidic sample environments with fixed sample materials, but there is usually still a flow that facilitates sample cooling). The third advantage is their small channel size, which allows high transmission of the beam. Except for very “hard” X-rays (>20 keV) used for X-ray radiography and tomography, beam path lengths of more than a few millimeters through liquids result in significant attenuation. Moreover, for “soft” X-rays (<2 keV) frequently used for spectroscopy and spectromicroscopy, it is difficult to measure samples of more than a few microns in thickness. Conversely, as stated earlier, depending on the energy of the incident X-rays, the thicker channel of a millifluidic device may optimize the produced signal relative to the amount of attenuation, at least in the case of X-ray scattering in a transmission geometry. Finally, the fourth advantage is the position-to-time conversion made possible along a flow channel.86 As will be discussed in greater detail below, this conversion enables the collection of data with high time resolution independent of instrumental hardware limitations.
2.3.2 Device material considerations. One of the first and most important considerations in designing a micro- or milli-fluidic device for X-ray analysis is selecting the device material(s). The previous review of Ghazal et al. covered these aspects well—and we encourage the reader to go there for more information12—but we will provide a brief overview here. As a general rule, any device material in the beam path should be as thin as possible and made from light elements. This minimizes the attenuation of the beam as well as whatever contribution the material might have to the overall X-ray signal, whether absorption, scattering, or fluorescence. Due to fabrication, chemical, thermal, and other practical constraints, this is not always possible. Thus, care must be taken to find a material that at least does not overly absorb, scatter, or fluoresce in a region of interest for your sample (N.B., at very hard X-ray energies, these restrictions are often relaxed). When a desired device material is not ideal for the X-ray technique of choice, it is also possible to design so called “windows” in your device made from a material with suitable X-ray characteristics. However, the integration of windows can present its own design challenges and provide additional locations for device failure (e.g., leakage, fracture, or hydrodynamic instabilities/eddies).

Many different device materials and fabrication approaches have been reported, as will be seen in the following review sections. Here, we will only introduce some of the most common families of device materials. The first are silicon and glass-based devices. These traditional microfluidic materials have the advantage of being highly chemically and thermally stable, and they can be patterned with high resolution features. However, they usually require expensive cleanroom fabrication methods and can present issues with high X-ray absorption and scattering—especially glass—unless using a very thin device or very hard X-rays.85 Silicon nitride (SixNy) is also commonly used as a window material since Si chips containing ultrathin low-absorption SixNy membranes (≤1 μm) can be readily purchased.87,88 However, these membranes are fragile, can bow/bend under fluid or vacuum pressure,89 and can be expensive if non-standard or low tolerance membrane sizes are required. Conversely, simple millifluidic devices often utilize a thin-walled glass capillary (∼10–100 μm) as the main analysis section. This is a cheap option with a very good signal-to-noise ratio owing to the relative thinness of the capillary wall with respect to its inner diameter.

The second group are curable polymers, such as polydimethylsiloxane (PDMS). While cheaper and easier to fabricate than microfluidic silicon/glass options, PDMS has a significant X-ray absorption and scattering profile,90 and many groups have demonstrated alternatives with better performance including Norland Optical Adhesive (NOA81)91 and off-stoichiometry thiol-ene (OSTE).92 Thirdly, another common option is using a commercial polymer film, the most popular of which is the polyimide, Kapton®. This polymer can be bought cheaply in thin sheets (∼25–100 μm), and it has excellent thermal and mechanical stability, moderate chemical resistance, and excellent resistance to X-ray radiation.93 It is a very good all-around material for microfluidic X-ray analysis; however, it does present some X-ray scattering features at small angles, which can introduce noise in SAXS data.92

Fourthly, high pressure/temperature flow cells are often made from metals or metal alloys due to their high thermal stability and mechanical strength. These are usually integrated with windows made from SixNy or diamond for use at lower X-ray energies but may be used without windows for hard X-ray tomography. Finally, some newer device materials, e.g., graphene94 and monocrystalline quartz,95 have been utilized for microfluidic fixed-target serial crystallography, and may find use for X-ray devices in the physical sciences. The X-ray absorption, transmission, and scattering properties of common device materials are found in many of the papers cited above and throughout the review and are widely available in previous microfluidic X-ray literature. Additionally, several helpful online calculators exist for estimating these parameters, such as from the Advanced Photon Source (https://11bm.xray.aps.anl.gov/absorb/absorb.php), Lawrence Berkeley National Laboratory (https://henke.lbl.gov/optical_constants/atten2.html), and the National Institute for Standards & Technology (https://physics.nist.gov/PhysRefData/FFast/html/form.html). The material information for all papers reviewed below can be found in Tables 1–4.

2.3.3 Important temporal definitions. While the numerous micro- and milli-fluidic devices intended for X-ray scattering, spectroscopy, and imaging applications have a variety of different design considerations and fabrication methods, there are many attributes and important parameters common to all platforms. Crucially, any sample environment should be well characterized to determine the conditions experienced by the sample and to enable accurate comparison with other experiments. A full treatment of the physical parameters and descriptors of microfluidic devices is outside the scope of this review, and the reader is directed to other sources for further information.82,83,96,97 Here, as this review is focused on the utility of microfluidic devices as sample environments for in situ and operando analysis, we will concentrate primarily on device operational parameters related to time.

One of the most important parameters to determine for a sample environment is the experimental or reaction time (t) associated with each measurement. This entails having high control over and low uncertainty in assigning t = 0 and understanding how the reaction or process develops in space across the device, e.g., by fluid flow or heating. Microfluidic devices are advantageous in this context, because as mentioned above, for a flow device operated under steady-state conditions, the position along a flow channel can be converted into an effective time. However, even for microfluidic devices, the task is not so straightforward, and there are several parameters and other characteristic times that must be considered before the true reaction time and time resolution of an experiment can be determined. Further, there are varying definitions for these terms in use throughout the literature. We seek to provide some clarity and standardization to these terms below.

For this discussion, we will use the example of the micro/millifluidic flow reactor, as it is a common sample environment used for a range of techniques and applications (Fig. 2). The most fundamental time parameter to consider for such a device is the mixing time (tmix) between the molecules or reagents that initiate the reaction. This time is critical because if reactants begin mixing at t = 0, but for example, it takes five seconds for them to mix, then any measurement made before 5 s of reaction time will contain some unreacted species and any measurement made afterwards will contain a mixture of reaction times: those starting closer to t = 0 and those starting closer to t = 5 s. Here, we define tmix as the time it takes to fully mix reactants, i.e., the time required from the initial contact of reactants to achieve uniform composition across the flow (Fig. 2a). Others may utilize a specific mixing index for defining tmix, for example, 90% mixed.98 Clearly, many reactions may begin before full mixing is achieved,69 thus experiments should ideally have tmix ≪ reaction time to decouple the mixing and the reaction. Failing this, the mixing should be at least faster than the reaction kinetics of interest to minimize uncertainty and prevent some kinetics from being masked. Although this is less important for thermally- or photo-induced reactions, for example, where species can be mixed slowly before the reaction is initiated further downstream.


image file: d4lc00637b-f2.tif
Fig. 2 Examples of (a) continuous and (b) droplet microfluidic flow reactors illustrating some important definitions of time related to sample environments for operando X-ray analysis. Reactants A and B are mixed at a Y-junction. Based on a given steady flow rate, the characteristic lengths shown correspond to characteristic times, t. A simple diffusive mixer is shown in (a), but various other designs could be used in the mixing region to accelerate the mixing process, i.e., by exploiting inertial effects for chaotic mixing.98 Mixing lengths are not to scale.

Mixing times are generally determined through numerical simulations and/or flow experiments using a colorimetric or fluorimetric tracer in the place of chemical reactants.99 Due to small measurement uncertainties or uncertainties in the diffusion coefficients of reactant molecules, often the mixing time is reported as a conservative upper limit or even presented simply as the observation dead time (tdead, Fig. 2a). This term is inherited from the stopped-flow community and simply means the time between the initiation of a reaction (t = 0) and the first possible measurement time. This distinction between tmix and tdead stems from the physical separation between the mixing element and the analysis capillary. Depending on the design of a microfluidic device, it may also not be possible to observe the flow right at the point of full mixing, or conversely, a time point after tmix may be targeted intentionally to allow for a factor of safety in the mixing time.

The next important time parameter is the time resolution (tres, Fig. 2a). This term is also defined in different ways in the literature, where it is sometimes taken to be equivalent to the mixing time, tmix. The logic for this definition is that it would be impossible to achieve higher time resolution than the distribution of fluid age resulting from mixing. To some extent this is accurate, however, if mixing is fast, often the limiting factor to resolution is the time it takes for fluid to pass through the beam, i.e., the age distribution of fluid within the beam neglecting mixing time. This is determined by the beam size in the direction of the flow and the fluid velocity, and it is one of several reasons why microfocused X-ray beams are typically utilized for microfluidics experiments. Improving temporal resolution can be a real advantage for the analysis of rapid kinetics, which are not able to be followed with the achievable acquisition times of most current X-ray instrumentation alone. For example, in the case of a microfocused beam (typically ∼20 μm in size) and an average linear velocity of 0.1 m s−1 (corresponding to a flow rate of 1 μL s−1 in a channel with a cross-section of 100 μm × 100 μm), one can obtain a temporal resolution of around 0.2 ms—much shorter than usual acquisition times.

Related to tres is the average time interval between each measurement position, or the time step (tstep, Fig. 2a). Depending on the device design and operation and the position of the X-ray windows, the distance between each measurement position can vary, with the smallest tstep without overlap being the effective length of the beam along the flow channel. This is often considered to be the full-width at half maximum (FWHM) of the beam intensity. If measurement positions are only a beam length apart, then tstep is equal to tres. Such an arrangement provides the highest possible resolution of the reaction taking place along the flow channel, but depending on the reaction kinetics, having many positions so close together may not be useful and will add unnecessary time and complication to an experiment. Importantly, like for tres, tstep is related to only the beam size and the fluid velocity with respect to the beam and is completely independent of the X-ray exposure time and detector frame rate unlike for stopped-flow experiments. There is some uncertainty in the time step arising from the precision of the sample stage motors and Taylor dispersion, i.e., the fact that fluid at the center of the channel will travel faster than fluid near the walls,100 although these effects are not often explicitly considered in operando X-ray experiments.

Alternatively, it is possible to eliminate Taylor dispersion by using a segmented flow of droplets in an immiscible continuous phase, where liquid and solids contained within a droplet stay together along a channel, forming an independent microreactor (Fig. 2b).100 In this case, tres is not defined by the beam size, but rather by the spread of reaction times in the fluid composing a droplet (i.e., tmix), assuming all droplets to be uniformly and continuously mixed. However, when utilizing droplets, an additional sampling consideration must be made to ensure that the signal of interest from the droplets is not masked by noise from the continuous phase.101 In droplet microfluidics, high density fluorinated oils are often utilized in this role, and these scatter significant numbers of photons. Thus, to minimize or eliminate noise from the continuous phase, the data acquisition strategy must adopt a droplet size, droplet velocity, beam size, and frame rate combination such that the effective acquisition length (or acquisition time, tacq) is contained within a single droplet (Fig. 2b).92 This consideration is analogous to the Nyquist rate in analog-to-digital signal conversion.102 For this reason, it is even more important to use a microfocused X-ray beam when performing an experiment with a droplet microfluidic device. Taylor dispersion can also be minimized using hydrodynamic flow focusing to concentrate reactants into a narrow fluid jet. However, this is only effective at short reaction times when diffusion can be neglected103 and also requires a microfocused X-ray beam to isolate data from the concentrated jet.

Now that we have learned about some alternatives to micro- and milli-fluidics and briefly discussed some characteristics and definitions for fluidic sample environments, the following three sections will cover the use of these devices in X-ray scattering & diffraction, spectroscopy, and imaging, respectively. The goal of these sections is to present as comprehensive of a view as possible into the work done in these areas in order to serve as a reference and to present the varied ways researchers have addressed the challenges of performing in situ X-ray measurements. These sections will also contain a greater focus on applications by discussing the science enabled by each device. Thus, to present as clear of a record as possible, the majority of our analysis and perspectives on the field will be included at the end of the review.

3 Devices for X-ray scattering and diffraction

3.1 Brief theory and overview

In microfluidic and millifluidic synchrotron X-ray analysis to date, scattering- and diffraction-based methods have received the most attention. The phenomenon of X-ray scattering results from elastic and inelastic interactions between incoming X-ray photons and the electrons within a sample.104 Focusing on the more common methods that utilize elastic X-ray scattering (of which diffraction is a special case), these provide the user with information on the length-based properties of a sample, such as its structure, size, and shape, depending on the scattering angle at which photons are analyzed.104,105 This dependence is well described by the relation formulated by W. L. Bragg in 1912:
 
λ = 2d[thin space (1/6-em)]sin[thin space (1/6-em)]θ (1)
where λ is the wavelength of the incident X-rays, θ is the scattering angle, and d is the associated distance within the sample.104 Consequently, for so-called “hard” X-rays with energies >5 keV (or wavelengths <2.5 Å) used in most scattering and diffraction experiments, wide angles above ∼5–10° provide atomic-scale structural information and smaller angles provide information on nanoscale features.

This distinction has led to the development of two groups of techniques: those based on X-ray diffraction (XRD) and those based on small-angle X-ray scattering (SAXS). The term “wide-angle X-ray scattering” (WAXS) also appears in the literature and is equivalent to X-ray diffraction. “XRD” is a term used by crystallographers and engineers, especially with crystalline samples that produce sharp Bragg peaks, whereas “WAXS” is preferred by SAXS practitioners, especially with poorly crystalline or amorphous samples. While there are a variety of different experimental setups for performing SAXS and XRD/WAXS analysis, the most obvious difference between the two groups is the position of the detector. Wide angles are analyzed when the detector is close to the sample (normally within ∼5–50 cm), whereas small angles are more easily accessed when the detector is a meter to several meters away or more. By convention, XRD data are typically plotted as a function of 2θ and scattering data as a function of the scattering vector, q = (4π/λ)sin(θ). Micro- and milli-fluidic devices for performing both SAXS and XRD/WAXS analysis are discussed below (Table 1).

Table 1 Micro- and milli-fluidic devices for X-ray scattering & diffraction
X-ray technique(s) Device material(s) Fabrication and/or assembly method Window material, thickness Geometry, beam pathlength Sample(s) investigated Beamline, source, X-ray energy Beam size Acquisition mode, exposure time Mixing time, tmix Minimum time step, tstep Total residence time Ref.
SAXS PDMS Photolithography, soft lithography, cured and plasma-bonded PDMS, 2 × 19 μm Transmission, 52 μm SiO2 NPs ID13, ESRF, 12.47 keV 1.5 μm × 1.5 μm Single-shot, 1 s N/A 160 s N/A Merlin et al. (2011)106
SAXS Kel-F, stainless steel, mica Machining, threaded Mica, unknown Transmission, 750 μm Au NPs with various ligands BL 11.3.1, ALS, 11 keV 100 μm spot Single-shot, 2–5 min N/A N/A 2 ms McKenzie et al. (2010)107
BL 7.3.3, ALS, 10 keV 0.24 mm × 1 mm
SAXS PMMA, Kapton Unknown Kapton, unknown Transmission, 1 mm Ag NPs BL08B2, SPring-8, 12.4 keV 230 μm × 370 μm Single-shot, 100 s Unknown 0.18 ms ∼1–10 s Takesue et al. (2011)108
SAXS Device 1: PDMS/glass Device 1: photolithography, cured and plasma-bonded Device 1: glass, 10 μm Device 1: transmission, 0.3–4 mm Au NPs 7T-MPW-SAXS, BESSY II, 7.5–8 keV 50 μm × 300 μm ellipse Single-shot, 900 s Device 1: N/A Device 1: unknown Device 1: unknown Stehle et al. (2013)109
Device 2: glass capillaries and slide Device 2: pipette pulling, epoxy Device 2: glass, unknown Device 2: transmission, unknown Device 2: ∼1 s Device 2: <0.1 s Device 2: ∼3–5 s
SAXS OSTEMER 322, Kapton Photolithography, soft lithography, cured, cure-bonded Kapton, 2 × 25 μm Transmission, 300 μm Cerium oxalate SWING, SOLEIL, 12 keV 80 μm × 150 μm 200 ms Unknown 33.5 ms ∼6–7 s Rodríguez-Ruiz et al. (2018)110
SAXS Unknown Compression fittings Capillary of unknown material and wall thickness Transmission, 2 mm outer diameter capillary Pd NPs 1–5, SSRL, 15.5 keV 500 μm × 500 μm Unknown Unknown N/A Unknown Fong et al. (2021)111
SAXS Titanium, diamond, unknown O-ring material Machined, clamped Diamond, 2 × unknown Transmission, unknown Supercritical CO2 4–2, SSRL, 15 keV Unknown Multiframe, 50–60 × 5 s N/A N/A N/A Younes et al. (2023)112
GISAXS COC (TOPAS), glass slide Machined, thermally bonded, clamped COC, 2 × 500 μm Reflection, 1 mm channel width, 7.45 mm beam footprint Au NPs BW4, DORIS III, 8.98 keV 65 μm × 35 μm Single-shot, 60 min Unknown N/A N/A Moulin et al. (2008)113
GISAXS COC (TOPAS), polymer-coated glass slide Machined, thermally bonded, clamped COC, 2 × 500 μm Reflection, 1 mm channel width, beam footprint unknown Au NPs BW4, DORIS III, 8.98 keV 30 μm × 60 μm Single-shot, 200 s N/A N/A N/A Metwalli et al. (2009)114
GISAXS/GIWAXS Silicon, glass (Pyrex) Photolithography, reactive ion etching, anodic bonding Silicon, 10 μm Reflection, unknown channel width, 2.9 mm beam footprint CO oxidation on RuO2 NPs cSAXS, SLS, 11.2 keV 10 μm × 100 μm Unknown Unknown N/A 23 s Kehres et al. (2016)115
SAXS/WAXS Glass capillaries, aluminum tube Pipette pulling, compression fittings PET (Melinex), 2 × 250 μm Transmission, 3 mm 2,6-Dibromo-4-nitroaniline 16.1, SRS Daresbury, 8.8 keV Unknown Single-shot, 30–60 s 90 ms 20 ms ∼1–2 s Alison et al. (2003)116
XRD Silicon tube, hastelloy Machined, clamped Silicon tube, unknown Transmission, 1 mm CaCO3 on silicon X17B1, NSLS, 67 keV Unknown Single-shot, 120 s Unknown N/A Unknown Chen et al. (2007)117
XRD Silicon tube, hastelloy Machined, clamped Silicon tube, unknown Transmission, 1 mm CaCO3 with polymer additive on silicon X17B1, NSLS, 67 keV Unknown Single-shot, 120 s Unknown N/A Unknown Chen et al. (2009)118
XRD Stainless steel, hastelloy Machined, clamped Stainless steel tube, unknown Transmission, 2 mm BaSO4 with polymer additives on stainless steel X17B1, NSLS, 70 keV Unknown Single-shot, 120 s Unknown N/A 0.75 s Mavredaki et al. (2011)119
GIWAXS Acetal plastic, Kapton Machined, clamped Kapton, 2 × 125 μm Reflection, 15 mm channel width, 717 μm beam footprint, 55 μm penetration depth FeCO3 on X65 stainless steel I15, diamond, 40 keV 100 μm diameter Multiframe, 5 × 60 s Unknown N/A Unknown Burkle et al. (2016)120
SAXS/WAXS Silicon, glass (Pyrex) Photolithography, wet etching, anodic bonding Glass, 250 μm, silicon, 280 μm Transmission, 220 μm CaCO3 ID02, ESRF, 12.49 keV 50 μm × 50 μm Single-shot, 200 ms Unknown N/A Unknown Beuvier et al. (2015)121
Powder XRD PMMA, PTFE, Kapton, silicone UV laser cutting, clamped Kapton, 2 × 75 μm Transmission, 300 μm CaCO3 with nucleating agents ID13, ESRF, 13 keV 12 μm × 15 μm Multiframe, 1000 × 20 ms ∼1 s 4 s 140 s Levenstein et al. (2020)102
I11, diamond, 15 keV 200 μm × 200 μm Single-shot, 60–120 s ∼10 s 9 s 325 s
Powder XRD PMMA, PTFE, Kapton, silicone UV laser cutting, clamped Kapton, 2 × 75 μm Transmission, 300 μm CaCO3 with nucleating agents, Au NPs, iron oxide NPs ID13, ESRF, 13 keV 12 μm × 15 μm Multiframe, 1000 × 20 ms ∼1 s 4 s 140 s Levenstein et al. (2019)122
I22, diamond, 12.4 keV 80 μm × 320 μm Multiframe, 2000 × 10 ms
Powder XRD FEP, PTFE, Kapton tubes Custom machined fittings, silicone sealant, tubing coiling Kapton tube, 95 μm wall Transmission, 3.19 mm Urea: barbituric acid, carbamazepine I11, diamond, 15 keV 1 mm × 1 mm Multiframe, 51 × 100 ms Unknown 138–230 s 10.4–19.3 min Levenstein, Wayment et al. (2020)123
SAXS and WAXS OSTEMER 322 Photolithography, soft lithography, cured and laminated OSTEMER 322, 2 × 200 μm Transmission, 150 μm Au NPs, cerium oxalate SWING, SOLEIL, 12 and 16 keV 50 μm × 125 μm Multiframe, 100 × 50 ms ∼0.5 s 6 ms 30 s Lange et al. (2020)92
SAXS/WAXS PDMS/glass Photolithography, soft lithography, cured and plasma-bonded Fused silica/Kapton tube, 50 μm wall Transmission, 250 μm Iron oxide NPs I22, diamond, 12.4 keV 40 μm × 40 μm Multiframe, 1000 × 20 ms ∼1 s 20 s 130 s Radajewski et al. (2021)124
Single-crystal XRD Custom formulated RLV-1 resin 3D printing (DLP) 3D-printed resin, 1.3 mm Transmission, ∼100–250 μm CaSO4·2H2O, protein crystals ID30-A3, ESRF, 12.82 keV 30 μm × 50 μm Unknown ∼1.5 s N/A N/A van der Linden et al. (2020)125
ID30-B, ESRF, 12.41 keV 10 μm × 10 μm Unknown
SAXS and WAXS Glass capillary, PTFE tubing, polymer fittings Tube crimping, compression fittings, heat-shrink tubing Borosilicate capillary, 50 μm wall Transmission, 2 mm YVO4:Eu NPs SWING, SOLEIL, 15 keV 375 μm × 75 μm Multiframe, 10 ms frames for SAXS and 0.1 or 2 s frames for WAXS ∼250 ms N/A 400 ms (longer using stopped flow and peristaltic pump) Fleury et al. (2014)126
Powder XRD Glass capillary, PTFE tubing, polymer fittings Compression fittings, tubing coiling Quartz capillary, 200 μm wall Transmission, 1 mm Iron oxide NPs XRD1, Elettra, 12.4 keV Unknown Single-shot, 10 min ∼50 ms N/A 5–160 s (longer using semi-batch setup) Besenhard et al. (2020)127
SAXS and WAXS Glass capillary, PTFE tubing, polymer fittings Tube crimping, compression fittings, heat-shrink tubing Borosilicate capillary, 50 μm wall Transmission, 1.5 mm Cerium oxalate SWING, SOLEIL, 16 keV 375 μm × 75 μm Multiframe, 20 × 1 s for SAXS and 20 × 4 s for WAXS ∼250 ms N/A 250 ms (longer using stopped flow) Durelle et al. (2023)128
Total scattering/PDF Kapton tube, metal fittings Compression fittings Kapton tube, unknown Transmission, 2 mm ZIF-8 28-ID-2, NSLS-II, 67.86 keV Unknown “Rapid acquisition mode”, ∼1–10 s Unknown N/A 0.05–2 s Terban et al. (2018)129
Total scattering/PDF Glass capillary, Kapton tube, stainless steel fittings Compression fittings, 3D-printed adapters Glass capillary, 100 μm wall Transmission, 1.3 mm or 0.9 mm Al3+, FeS 11-ID-B, APS, 58.62 keV 0.5 mm × 0.5 mm Single-shot, 1, 10, and 100 s ≤0.1 s 0.01–0.1 s 0.1–10 s Beauvais et al. (2021)130
28-ID-1, NSLS-II, 74.46 keV Beauvais et al. (2022)131
Total scattering/PDF Carbon fiber capillary, polyethylene, rubber tubing Compression fittings Epoxy-aligned carbon fiber capillary, 0.67 mm Transmission, 1.83 mm Pt NP-coated CNTs and graphene, Fe–Ni layered double hydroxide NPs, LiCoO2 11-ID-B, APS, 58.65 keV Unknown Unknown N/A N/A N/A Young et al. (2017)132
11-ID-C, APS, 105 keV
6-ID-D, APS, 100 keV
Total scattering/PDF and GIWAXS Kapton, VeroClear-RGD810 resin, porous glass array, electrode materials Epoxy, atomic layer deposition of electrodes, compression fittings Kapton, 2 × 25 μm Transmission, 2.4 mm Amorphous cobalt oxide thin films 11-ID-B, APS, 58.7 keV 300 μm × 500 μm Single-shot, 2–5 min N/A N/A N/A Kwon et al. (2019)133
6-ID-D, APS, 100.3 keV 300 μm × 500 μm
11-ID-D, APS, 23 keV 15 μm vertical


3.2 SAXS

3.2.1 Early work. Microfluidic SAXS analysis began around the year 2000, when it was utilized to study protein folding in response to chemical stimuli.134,135 To our knowledge, the first microfluidic SAXS experiments on inorganic samples did not take place until close to a decade later. Merlin et al. used a polydimethylsiloxane (PDMS) device to study the concentration of a suspension of silica nanoparticles (NPs) by solvent evaporation.106 The main channel of the device was made within a thin PDMS membrane (∼20 μm wall thickness) to enable the transmission of X-rays. The gas permeability of this thin PDMS layer was also exploited to create a steady-state evaporation-driven flow of water, which slowly concentrated the NPs from the end of the microchannel inwards, until they were converted into a solid-state colloidal crystal at a NP volume fraction of 30–40%. By using a high flux, microfocused beam, the researchers were able to perform the study with high spatial and temporal resolution, scanning the beam along the channel in a grid with 50 μm steps and 1 s X-ray exposure time per step.

Around the same period, several millifluidic studies were also performed. McKenzie et al. designed a custom flow-cell to enable simultaneous in situ SAXS and ultraviolet-visible (UV-vis) spectroscopy and validated it by determining the size distribution of pre-made suspensions of reference Au NPs (Fig. 3a).107 The ability to perform both in situ and ex situ UV-vis allowed for quality control and the comparison of in situ SAXS data to subsequent ex situ transmission electron microscopy (TEM) of the different Au NP standards after surface-deposition, which would be especially important when characterizing experimental samples. Takesue et al. performed operando SAXS analysis of Ag NP synthesis using a poly(methyl methacrylate) (PMMA) device with Kapton X-ray windows (Fig. 3b).108 By utilizing a very high flow rate (120 mL min−1), the authors obtained a turbulent flow, which facilitated rapid mixing of reactants and also sub-ms time resolution through the vertical movement of the device in the beam. In this case, the continuous flow of solution permitted long X-ray exposures (>1 min) at each channel position to obtain good scattering statistics of dilute intermediate species without sacrificing time resolution.


image file: d4lc00637b-f3.tif
Fig. 3 Devices for SAXS analysis. (a) Millifluidic flow-cell for simultaneous SAXS and UV-vis of nanoparticle solutions. Inset: Example 1D scattering pattern of AuNP suspension (adapted with permission from McKenzie et al., 2010; Copyright 2010 American Chemical Society)107 (b) Experimental setup for sub-ms synchrotron SAXS measurements of the early stages of AgNP synthesis (reprinted with permission from Takesue et al., 2011; Copyright 2011 American Chemical Society).108 (c) Continuous flow microfluidic device for the study of cerium oxalate precipitation (top left: photo of the device; top right: measurement geometry; bottom: illustration of the mixing configuration with the water buffer flow; adapted with permission from Rodríguez-Ruiz et al. 2018; Royal Society of Chemistry).110 (d) Example of a microfluidic GISAXS experiment (reprinted with permission from Metwalli et al., 2009; Copyright 2009 American Chemical Society).114
3.2.2 Mitigation of device fouling. In many precipitation and synthesis scenarios, the continuous flow of solution can result in the build-up of products on channel walls.136,137 This can convolute or degrade the X-ray signal, change the chemical environment and modify reaction products, and potentially lead to device failure.102,138 For this reason, Stehle et al. performed the first SAXS analysis of nanoparticle synthesis using droplet microfluidics.109 They utilized two devices: one in which a pre-made solution of Au NPs was emulsified in a PDMS chip and injected into a glass capillary for analysis and another device made from nested glass capillaries where the synthesis and analysis were both performed in situ. They were able to obtain a SAXS signal from the Au NPs synthesized in situ, however, the long exposure times employed (900 s) meant that the acquired scattering curves contained a mixture of scattering from many droplets and the surrounding continuous oil phase. We will see in the next subsection different routes for isolating droplet scattering from that of the rest of the flow to obtain a higher signal-to-noise ratio.

Another method for limiting precipitation on channel walls without producing droplets is by introducing a strong ‘buffer’ flow of water between reactant streams. This delays their contact and slows down their mixing to prevent a sudden precipitation event that could instantly clog a device. Such a method was employed by Rodríguez-Ruiz et al. in an OSTEMER-Kapton microfluidic device to study the precipitation of the highly insoluble, rare earth mineral, cerium oxalate (Fig. 3c).110 Using a flow of water >10 times faster than their reactant flows (also resulting in a >10 times reactant dilution), they were able to successfully acquire scattering curves from within the first second of the reaction. However, the slowed mixing resulting from the water flow meant that it was not possible to analyze reaction times <0.2 s due to inconsistent background signal.

3.2.3 Machine learning. An exciting recent application of in situ SAXS comes from Fong et al., who used a millifluidic continuous flow reactor and machine learning to optimize the synthesis of Pd nanoparticles in real-time.111 Their setup consisted of pumps containing the reactants, surfactants, and solvents coupled to a mixing manifold and heated capillary tube. SAXS data acquired at the outlet of the tube were automatically processed by a control computer, which could then adjust the flow rates and reactor temperature in a closed feedback loop. After performing a short grid search to roughly map a reaction parameter space, the system was able to rapidly converge on conditions to synthesize nanoparticles with targeted size, concentration, and yield using Bayesian optimization (BO). Similarly, Younes et al. used BO to efficiently determine the phase diagram of supercritical CO2 using SAXS and a high-pressure microfluidic cell.112 Correlation lengths were extracted from SAXS curves obtained at different temperature and pressure combinations, and the BO algorithm was shown to accurately predict the correlation length maximum for each temperature after measuring at only 5–10 different pressures. This required 3–6 fewer measurements with the microfluidic device than when using a more traditional statistical approach (the bisection method) and could be used in the future to eliminate unnecessary data collection.
3.2.4 GISAXS. Standard SAXS measurements made using a transmission geometry are good for obtaining information from bulk samples and liquids, but they are less sensitive to surface features.139 Therefore, there has also been interest in performing grazing incidence SAXS (GISAXS) in a reflection geometry to study interfacial phenomena occurring within microfluidic devices. Moulin et al. developed a device for this purpose made from cyclic olefin copolymer (COC, TOPAS®) that could be reversibly clamped to various substrates of interest (Fig. 3d).113 They confirmed its suitability for GISAXS by studying the interaction between a flow of Au NPs and a glass surface. Later, the team followed up this study by using the same device to monitor the templated growth of Au nanowires on a polymer substrate.114 Similarly, Kehres et al. designed a silicon-based device for studying gas–solid interactions in catalysis.115 Prior to device bonding, a catalyst layer is deposited, and then catalytic reactions at the interface can be monitored using different gas mixtures, temperatures, and pressures. Their device was also configured to work with inline mass spectrometry while simultaneously collecting grazing incidence WAXS (GIWAXS) data.

3.3 XRD/WAXS

3.3.1 Millifluidics. While in situ microfluidic XRD/WAXS analysis is a more recent trend than microfluidic SAXS (excluding for structural biology), several groups implemented millifluidic XRD/WAXS quite early, and often combined with SAXS. Alison et al. utilized a glass capillary-based plug-flow reactor to investigate the crystallization of 2,6-dibromo-4-nitroaniline, a model compound for studying crystallization mechanisms.116 Using a high flow rate and a turbulent ‘teardrop’ mixer, they observed the formation of non-crystalline particles by SAXS during the first second of precipitation and prior to the appearance of crystalline diffraction observed by WAXS. This combination of techniques thus allowed them to evidence potential crystallization pathways that do not follow single-step classical nucleation theory, which does not account for the formation of non-crystalline intermediate phases. Likewise, Chen et al. developed a millifluidic flow cell for studying scale formation from salt brines on different materials (including silicon and stainless steel).117 Beginning with a study of the effect of temperature on CaCO3 scaling, the authors went on to study both BaSO4 and CaCO3 scaling in the presence of anti-scaling agents.118,119 Later, the same team developed an updated flow-cell that could perform GIWAXS studies of steel corrosion combined with simultaneous electrochemical monitoring (Fig. 4a).120
image file: d4lc00637b-f4.tif
Fig. 4 Devices for WAXS/XRD analysis. (a) An experimental setup for millifluidic GIWAXS and electrochemical measurement of steel corrosion (reprinted from Burkle et al. 2016 with the permission of AIP Publishing).120 (b) A hybrid silicon-glass microfluidic device for continuous flow study of CaCO3 crystallization (reprinted with permission from Beuvier et al. 2015; Royal Society of Chemistry).121 (c) Comparison of CaCO3 crystallization in microfluidic devices in continuous flow (top left) versus under conditions optimized to prevent scaling (top right). An illustration of the distance-to-time conversion enabled when scaling is prevented (bottom) (adapted with permission from Levenstein et al., 2020; Royal Society of Chemistry).102 (d) A workflow for SAXS/WAXS frame selection and background subtraction for droplet microfluidics (reprinted with permission from Radajewski et al., 2021; Royal Society of Chemistry).124
3.3.2 Microfluidics. More recently, researchers have miniaturized these WAXS and SAXS/WAXS studies further by utilizing microfluidic devices. Beuvier et al. studied the precipitation of CaCO3 by SAXS/WAXS using a hybrid silicon-glass device (Fig. 4b),121 where the single-crystalline silicon layer provided a very low SAXS background and the glass layer permitted visual monitoring of the flow. WAXS enabled the identification of the structure/polymorph of the CaCO3 crystals formed, and the presence of streak-like crystal truncation rods in the SAXS patterns enabled the determination of their micron-scale size. Similar to the above studies of mineral scaling, the authors observed the rapid growth of CaCO3 on device surfaces starting from the point of reactant mixing. However, as discussed in the previous section, for many applications, such surface growth is undesirable, making it more difficult to study processes occurring in the bulk flow. Levenstein et al. also performed a WAXS study of CaCO3 growth in a microfluidic device in order to test several different flow configurations and determine those best for minimizing device scaling.102 They identified that a combination of water buffer between reactants, hydrophobic surface treatment, segmented water-in-fluorinated oil flow, and triblock co-polymer surfactant facilitated studies of CaCO3 precipitation for ∼1 hour without significant crystallization on surfaces (Fig. 4c). Importantly, this enabled the crystallization process to be studied as a function of a droplet's position along the channel with time, where the rotation of crystals within the droplets additionally allowed diffraction to be collected from a range of scattering angles θ.

In order to isolate scattering from the droplets from that of the surrounding fluorinated oil phase, the authors implemented a multiframe data acquisition and processing approach first used for bioSAXS.101 Briefly, short 20 ms frames were captured at a rate of 50 Hz and WAXS frames containing the characteristic scattering of the oil phase were identified and discarded. The remaining frames were then summed to obtain a good signal-to-noise ratio at each position.122 This WAXS-based technique was later implemented in a millifluidic flow system,123 and a similar technique was also performed by Lange et al., who utilized SAXS frames to distinguish between the water and oil phases.92 More recently, Radajewski et al. presented an innovative data processing technique combining both WAXS and SAXS frame selection to isolate not only droplets, but also the sections of droplets with the highest concentration of sample for subsequent data treatment (Fig. 4d).124 Alternatively, for studies not requiring operando measurements, van der Linden et al. developed a 3D-printed device for storing and measuring samples contained with isolated, static droplets to avoid signal from the oil phase.125

3.3.3 Plug-and-play devices. Despite this progress in microfluidic SAXS/WAXS studies, a growing trend is the use of simpler ‘plug-and-play’ millifluidic devices and set-ups that lower the fabrication, cost, and time barriers to performing in situ flow experiments. These often make use of commercially available compression fittings (e.g., Upchurch/IDEX, Swagelok) and/or mixers, the outlet of which is connected to a thin-walled glass capillary for performing SAXS and/or WAXS analysis. For example, Fleury et al. made use of a simple home-made turbulent mixer140 to study the synthesis of luminescent Eu-doped YVO4 nanoparticles.126 Using combined SAXS, WAXS and fluorescence spectroscopy, they observed that the final fluorescent nanoparticles were formed from non-fluorescent amorphous aggregates, the size of which defined the size of the product crystallites. Similarly, Besenhard et al. studied the precipitation of iron oxide nanoparticles by XRD using commercial mixers combined with custom temperature controlled modules of varying lengths to adjust the age of the material probed in the beam (Fig. 5a).127 They additionally tested mixers with different internal diameters, and thereby different Reynolds numbers, and showed that they produced similarly sized nanoparticles with production that was consistent over >30 min of operation. Recently, Durelle et al. used a simple device to compare in situ SAXS and WAXS measurements of crystal nucleation rates to traditional incubation-counting methods used in chemical engineering and found that traditional methods could result in underestimates of several orders of magnitude.128 This illustrates clearly the power of even simple millifluidic continuous flow devices for in situ X-ray scattering. Many other studies utilizing such devices exist and cannot all be covered here.141–143
image file: d4lc00637b-f5.tif
Fig. 5 Simple ‘plug-and-play’ millifluidic devices for XRD/WAXS/PDF analysis. (a) A setup for in situ SAXS/WAXS utilizing commercial mixers and custom temperature-controlled sections for time-delay (reprinted from Besenhard et al., 2020; with permission from Elsevier).127 (b) A Norby-style device with active mixing for X-ray PDF analysis (adapted with permission from Beauvais et al. 2022; IUCr).131

3.4 Total scattering/PDF

In this final scattering section, we will cover recent efforts in millifluidic X-ray total scattering, used for performing pair distribution function (PDF) analysis of atomic to nanoscale correlations within materials. These measurements are usually performed at high energy beamlines to incorporate scattering from small length scales since wavelength and energy are inversely proportional and wavelength and the probed length scale are directly proportional at fixed θ (see eqn (1)). By utilizing crystalline Bragg diffraction and diffuse scattering at large angles normally neglected during XRD analysis (i.e., the “total” scattering pattern), the PDF technique provides information on not only long-range crystalline order within a sample, but also short-range correlations within even amorphous materials.144 These techniques are additionally well suited for in situ analysis because the high energies impart more penetration power and allow the use of thicker and denser samples and sample environments.

Terban et al. performed in situ X-ray PDF analysis of the synthesis of the zeolitic imidazolate framework, ZIF-8,129 using a simple millifluidic continuous flow device consisting of a Kapton tube, a metallic frame and commercial fittings mounted on a goniometer head in the same way as a Norby cell. With this setup, they were able observe the formation of long-lived solution species and amorphous solid phases during the synthesis of this model metal–organic framework. Similarly, Beauvais et al. developed a millifluidic device resembling a Norby design but comprising an active vibration mixing element (Fig. 5b).131 They tested a variety of injection capillary sizes, types and materials, including glass and Kapton, and converged on a design that fully mixed reactants in less than a few hundred milliseconds. The authors validated the system by studying the hydrolysis of Al3+ and then went on to study the formation of FeS by a ligand-exchange reaction, demonstrating the presence of previously unknown intermediate phases in the form of nanosheets.130

A few groups have also developed innovative cells that enable in situ X-ray PDF measurements of electrochemical processes. For example, Young et al. designed a millifluidic electrochemical cell with an epoxy-aligned carbon fiber capillary serving as both the working electrode and the X-ray window.132 A recirculating flow of electrolyte solution could be applied through the capillary, and various samples relevant to electrocatalysis and batteries could be loaded and measured during cycling, including Pt nanoparticle-coated carbon nanotubes and LiCoO2 powder. Kwon et al. developed their own electrochemical cell consisting of a 3D porous glass capillary-array (GCA) sitting in an electrolyte reservoir with Kapton walls to facilitate X-ray analysis.133 The GCA array was coated by gold and either indium tin oxide (ITO) or indium zinc oxide (IZO) to serve as the working electrode, and fresh electrolyte was pumped through the GCA pores from below the reservoir using a syringe pump during cycling.

4 Devices for X-ray spectroscopy

4.1 Brief theory and overview

X-ray spectroscopy techniques are utilized to obtain chemical information such as elemental composition, oxidation state, and coordination number. This section will be focused primarily on X-ray absorption spectroscopy (XAS), which is one of the main groups of X-ray spectroscopic techniques and the one utilized in most previous micro- and milli-fluidic studies. Another major type of X-ray spectroscopy is X-ray photoelectron spectroscopy (XPS), but to our knowledge, in situ XPS has been performed exclusively in free liquid jets,145 likely due to the poor transmission of electrons exiting through device windows. There are two primary acquisition modes for XAS analysis: transmission and fluorescence. The most common, transmission XAS, is based on measuring the incident beam flux (I0) and the flux of the beam transmitted through the sample (I), which are related by the Beer–Lambert law:
 
I = I0·eμ(λ)x (2)
where μ(λ) is the wavelength-dependent linear attenuation coefficient of the sample and x is the sample thickness.146 XAS spectra are collected by scanning the wavelength of the incident X-ray beam across an absorption edge of an element of interest in the sample, where the shape and position of the edge and post-edge oscillations are used to determine the chemical speciation of the target element. The location of the collected spectra in relation to the absorption edge subdivides XAS into different techniques based on where the data are collected, such as X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopy for the edge and post-edge regions, respectively.146

XAS is performed in fluorescence mode by monitoring the total yield of secondary X-rays at 90° from the incident beam (normally 45° from the sample) rather than the transmitted flux, which is helpful when using thick or highly dilute samples.146,147 For both modes, different acquisition strategies using polychromatic radiation and position/energy-resolved detectors can also be employed to more rapidly record full spectra without scanning the energy of the incident beam.148–150 The closely related technique of X-ray fluorescence (XRF) has also been performed with microfluidic devices,151,152 although primarily for 2D elemental mapping and laboratory analysis, and will thus be covered in sections 5 and 6, respectively. Further details on the micro- and milli-fluidic devices for X-ray spectroscopy discussed below can be found in Table 2.

Table 2 Micro- and milli-fluidic devices for X-ray spectroscopy
X-ray technique(s) Device material(s) Fabrication and/or assembly method Window material, thickness Mode, beam pathlength Sample(s) investigated Beamline, source, X-ray energy Beam size Acquisition mode, exposure time Mixing time, tmix Minimum time step, tstep Total residence time Ref.
XRF Fused silica capillary, polyethylene tubing Interference fit Polyethylene, 193 μm wall Fluorescence, 580 μm Co, Cu, and Zn solutions X-26A, NSLS, 10 keV 30 μm × 40 μm Multiframe, 1 s exposure, 0.25 Hz N/A N/A N/A Ringo et al. (1999)151
XRF Fused silica capillary, polyethylene tubing Interference fit Polyethylene, 355 μm wall Fluorescence, 380 μm Fe, Co, Cu, and Zn solutions X-26A, NSLS, 10 keV 40 μm × 40 μm Multiframe, 1 s exposure N/A N/A N/A Mann et al. (2000)152
XANES and EXAFS Ti-6Al-4V alloy, Pt/Ir alloy, diamond (type 1B) Machining, press fit, clamping Diamond, 2 × 1 mm Transmission, ∼1–100 mm Tungstate solutions ID-20, APS, W L3-edge (10.21 keV) 1.5 mm diameter Unknown N/A N/A N/A Hoffmann et al. (2000)153
Diamond, 2 × 250 μm Transmission, ∼2.5 mm Chromate solutions X-19A, NSLS, W L3-edge (10.21 keV) 2 mm diameter Hoffmann et al. (2001)154
ID-20, APS, Cr K-edge (5.99 keV) Unknown
XANES and EXAFS Silicon, glass Deep reactive ion etching, anodic bonding Silicon, unknown Fluorescence, ≥250 μm Dehydrogenation of methanol on Ag 16.5, SRS Daresbury, Ag K-edge (25.5 keV) 0.4 mm × 10 mm Single-shot, 40 min N/A N/A ∼10 ms gas dwell time Sankar et al. (2007)155
XANES and μXRF Silicon, glass, silicon nitride, SU-8 Photolithography, wet etching, reactive ion etching, anodic bonding Silicon nitride, 1 μm, SU-8, 1 μm Fluorescence, 570 μm CdSe → Ag2Se nanocrystals 10.3.2, ALS, Se K-edge (12.66 keV) 16 μm × 7 μm Multiframe, 4 × >100 s ∼14 ms ∼8–16 ms ∼50–100 ms Chan et al. (2007)156
XANES Silicon, glass Deep reactive ion etching, fusion bonding Silicon, 300 μm, glass, 200 μm Fluorescence, ≥300 μm Au NPs P06, Petra III, Au L3-edge (11.92 keV) Unknown Unknown <1 ms ∼1 μs 20 ms Hofmann et al. (2016)157
XANES Silicon, glass Deep reactive ion etching, fusion bonding Silicon, 300 μm, glass, 200 μm Fluorescence, ≥300 μm Au NPs SuperXAS, SLS, Au L3-edge (11.92 keV) 150 μm × 100 μm Single-shot, 4 min <2 ms ∼1 μs 20 ms Tofighi et al. (2017)158
XANES and EXAFS Silicon, borosilicate glass Deep reactive ion etching, anodic bonding Glass, 2 × 175 μm Transmission and fluorescence, 742 μm Fe, Br, and Pb salt solutions Balder, MAX IV, Fe K-edge (7.11 keV), Pb L3-edge (13.04 keV), Br K-edge (13.47 keV) 50 μm × 50 μm Multiframe, 10 × 100 s for EXAFS and 4–6 × 25 s for XANES N/A N/A N/A Micheal Raj et al. (2021)159
XANES/EXAFS Unknown Unknown Unknown Fluorescence, unknown Pt NPs I18, diamond, Pt L3-edge (11.56 keV) 400 μm × ∼250 μm Unknown Unknown Unknown 37.3 min Britto et al. (2023)160
XANES PMMA Hot embossing PMMA, unknown Fluorescence, ≥500 μm Co NPs XMP, CAMD, Co K-edge (7.78 keV) 50 μm × 80 μm Single-shot, 5–7.5 min Unknown ∼2 ms 50 s Zinoveva et al. (2007)161
XANES and EXAFS Unknown Unknown Kapton, unknown Fluorescence, ≥500 μm CdSe NPs BL13B1, PF, 12.6–12.7 keV 1 mm × 0.5 mm Unknown Unknown 47–118 ms 30 s Oyanagi et al. (2011)162
NW2, PF-AR, 12.6–12.7 keV
XANES and EXAFS PVC, brass, graphite Machining, clamping, pressure fittings, conductive epoxy Kapton, unknown, graphite, 500 μm Transmission, ∼400 μm (∼200 μm electrolyte and ∼200 μm sample) Iron and iron oxide phases ID24, ESRF, Fe K-edge (7.11 keV) 50 μm × 100 μm Multiframe, ms exposures N/A N/A Unknown Monnier et al. (2008)163
Fluorescence, 282 μm BM30b, ESRF, Fe K-edge (7.11 keV) 30 μm × 150 μm Multiframe, 7 min for XANES, and 3 × 30 min for XANES + EXAFS Monnier et al. (2014)164
DIFFABS, SOLEIL, K-edge (7.11 keV) 300 μm × 300 μm Multiframe, unknown
XANES and EXAFS PET Commercially purchased PET, unknown Fluorescence, ≥150 μm Au nanostructures 10-ID, APS, Au L3-edge (11.92 keV) 50 μm × 50 μm Unknown Unknown 5.4 ms (neglecting mixing time) ∼25 s Krishna et al. (2013)165
WDCM, CAMD, Au L3-edge (11.92 keV)
XANES and EXAFS PETG filament 3D printing (FDM) 3D-printed polymer, unknown Transmission, 80 mm Pd NPs STM, Kurchatov, Pd K-edge (24.35 keV) 0.7 mm × 0.7 mm Single-shot, 10 min Unknown ∼2.7 min ∼17.6 min Dobrovolskaya et al. (2023)166
EXAFS Ti-6Al-4V alloy, diamond (type IIa) Machining, epoxy, clamping, Poulter seal Diamond, 2 × 25 μm Transmission, 150 μm CaCl2 solution 20-BM, APS, Ca K-edge (4.04 keV) 200 μm × 180 μm Multiframe, 3 × 20 min N/A N/A ∼15 min Fulton et al. (2004)167
XAS and XES PTFE, silicon, viton O-ring Machined, clamped Silicon nitride, 100 nm silicon carbide, 150 nm Fluorescence, ≥500 μm H2O, D2O 8.0.1, ALS, 550 eV Unknown Unknown N/A N/A Unknown Fuchs et al. (2008)168
XANES Silicon, PTFE, unknown O-ring and housing material Machined, clamped Silicon nitride, 2 × 100 nm Transmission, 100–800 nm Water BL3U, UVSOR-II O K-edge (532 eV) 200 μm × 200 μm Single-shot, ∼13 min N/A N/A ∼1 s Nagasaka et al. (2010)169
XANES Silicon, stainless steel, gold, unknown O-ring material Machined, clamped Silicon nitride, 2 × 100 nm Transmission, 250 nm Water, CoCl2 solution, methanol–water mixture PM3, BESSY II, O K-edge (532 eV) and Co L3-edge (778.6 eV) 100 μm diameter Unknown N/A N/A Unknown Schreck et al. (2011)170
XANES and μXRF Silicon, PDMS Wet etching, photolithography, soft lithography, plasma bonding Silicon nitride, 450 nm Fluorescence, 57 μm CaCO3 Phoenix, SLS, Ca K-edge (4.04 keV) 50 μm × 75 μm Multiframe, 10–15 × 8 min <10 ms ∼1 ms ∼5 s Probst et al. (2021)171
XANES and μXRF Silicon, PDMS or glass Wet etching, deep reactive ion etching, photolithography, soft lithography, plasma bonding Silicon nitride, 120 nm Fluorescence, 127 μm Ca2+ ions and EDTA Phoenix, SLS, Ca K-edge (4.04 keV) 3 μm × 3 μm Multiframe, ∼100 min total <2 ms 0.365 ms 2.7 s Brenker et al. (2022)172
XANES/EXAFS Stainless steel, graphite, quartz wool Machined, clamped Graphite, 300 μm CaF2/glue, 300 μm Transmission, 2 mm Pt/Al2O3 catalyst SuperSAXS, SLS, Pt L3-edge (11.56 keV) 100 μm × 100 μm Single-shot, 1 s N/A N/A ∼5 s Chiarello et al. (2014)173
XANES/EXAFS Aluminum, CaF2 Machined, pressure fittings Al, 2 × 250 μm Transmission, 5 mm Pd/Al2O3 catalyst ID12-EDE, diamond, Pd K-edge (24.35 keV) 500 μm × 150 μm Multiframe, 200 × 4.8 ms for EXAFS, single-shot 4.8 ms for XANES N/A N/A Unknown Dann et al. (2019)174
XANES/EXAFS Silicon, glass Photolithography, deep reactive ion etching, anodic bonding Si, 2 × 250 μm Transmission, 3 mm Pd/Al2O3 catalyst B18, diamond, Pd K-edge (24.35 keV) 200 μm × 100 μm Single-shot, 180 s N/A N/A Unknown Venezia et al. (2020)175
XANES and XES Fused silica, Kapton Laser-selective wet etching Kapton tube, 27 μm wall thickness Fluorescence, 510 μm Ferricyanide and ascorbic acid 6-2b, SSRL, Fe K-edge (7.11 keV) 418 μm horizontal Single-shot, 45 min <1 ms <1 ms 157 ms Huyke et al. (2021)176
XANES/EXAFS and SAXS Silicon, glass (Pyrex) Photolithography, deep reactive ion etching, anodic bonding XAS: silicon, ∼50 μm XAS: fluorescence, ∼382 μm Pb NPs XAS: X18B, NSLS, and 10-ID-B, APS, Pd K-edge (24.35 keV) XAS: X18B, 0.2 mm × 3 mm, 10-ID-B, 0.5 mm × 0.5 mm Multiframe, 5 × 20 min Unknown 30 s 95 min Karim et al. (2015)177
SAXS: silicon, ∼50 μm + second layer of unknown thickness SAXS: transmission, ∼270 μm SAXS: 12-ID-C, APS, 18 keV and 12-ID-B, APS, 12 keV SAXS: unknown
XANES/EXAFS and anomalous SAXS PEEK, Kapton (gold-coated) Machined, clamped Kapton, 2 × 50 μm Transmission, 10–50 μm of catalyst, 2 mm of electrolyte Pt/IrO2–TiO2 electrocatalyst, HClO4 electrolyte XAS: SuperXAS, SLS, Ir L3-edge (11.22 keV) XAS: 100 μm × 100 μm XAS: multiframe, 120 × 500 ms N/A N/A N/A Binninger et al. (2016)178
SAXS: cSAXS, SLS, four energies near the Pt L3-edge (11.56 keV) SAXS: unknown SAXS: unknown
XANES and SAXS OSTEMER 322, Kapton Photolithography, PDMS injection molding, cured, cure-bonded Kapton, 2 × 75 μm XANES: transmission, 1.9 mm Au NPs XAS: SuperXAS, SLS, Au L3-edge (11.92 keV) XAS: 20 μm × 20 μm XAS: unknown <0.3 ms ∼100 ms XAS: 30 s Ramamoorthy et al. (2024)179
SAXS: transmission, 370 μm SAXS: cSAXS, SLS, 11.5 keV SAXS: 20 μm × 50 μm SAXS: unknown SAXS: 0.3 s
XPCS Aluminum, Kapton Machining, clamping Kapton, 2 × unknown Transmission, 1 mm Latex NPs ID10A, ESRF, 8 keV 10 μm × 10 μm Unknown N/A N/A N/A Busch et al. (2008)180
XPCS Quartz capillary tube Compression fittings Quartz capillary, unknown wall thickness Transmission, 0.98 mm PMMA NPs ID10A, ESRF, 8 keV 10 μm × 10 μm Unknown N/A N/A N/A Fluerasu et al. (2008)181
XPCS Kapton tube Compression fittings Kapton tube, 100 μm wall Transmission, 1.32 mm SiO2 NPs ID10A, ESRF, 8 keV 10 μm × 10 μm Unknown N/A N/A N/A Fluerasu et al. (2010)182
XPCS NOA 81, polystyrene Photolithography, soft lithography, UV-curing, cure-bonding Polystyrene, 2 × 50 μm Transmission, 200 μm SiO2 NPs P10, Petra III, 8.05 keV 5 μm × 5 μm Multiframe, 5000 × 3.33 ms N/A N/A N/A Urbani et al. (2016)183
Heterodyne XPCS Copper, Kapton Machining, epoxy, clamping Kapton, 2 × unknown Transmission, 0.69 or 0.8 mm SiO2 NPs 8-ID-I, APS, unknown 5 μm × 20 μm Multiframe, unknown × 1.25 or 16.67 ms N/A N/A N/A Lhermitte et al. (2017)184
Heterodyne XPCS and XAM PEEK, stainless steel Compression fittings, epoxy PEEK, 2 × 2 mm Transmission, 1 mm Li/PEO–LiTFSI/Li battery cell 8-ID-I, APS, 11 keV 15 μm × 15 μm Multiframe, 6000 × 60 ms N/A N/A N/A Steinrück et al. (2020)185
ID10, ESRF, 8.1 keV 10 μm × 15 μm Unknown
sp-XPCS, XPXP, SAXS, WAXS Ti-6Al-4V alloy, diamond (type IIa) Machining, epoxy, clamping, Poulter seal Diamond, 2 × 100 μm Transmission, ∼0.4–1 mm Supercritical H2O sp-XPCS: LCLS, 9.5 keV sp-XPCS: 3 μm diameter sp-XPCS: multiframe, 105 × 8.33 ms N/A N/A N/A Muhunthan et al. (2024)186
XPXP: SACLA, unknown XPXP: unknown XPXP: unknown
SAXS: BL4-2, SSRL, 15 keV SAXS: unknown SAXS: unknown


4.2 XAS

4.2.1 Hard X-rays. Unlike with scattering methods, microfluidic XAS was developed and has been primarily conducted for the physical rather than biological sciences. Early work in micro- and milli-fluidic XAS was focused on devices designed for analyzing heavy elements at hard X-ray energies. The first devices, such as from Hoffman et al., were made from corrosion-resistant metal alloys and diamond windows for studying high pressure phenomena.153,154 Subsequent devices were made from silicon and glass using more conventional microfluidic fabrication processes and applied to catalysis and nanoparticle synthesis. Sankar et al. utilized such a device to study the dehydrogenation of methanol on an silver catalyst with XANES and EXAFS.155 Using fluorescence mode acquisition at the Ag K-edge (25.5 keV), the authors evidenced a cyclical process of oxidation/reduction of the metal catalyst involved in the conversion of methanol to formaldehyde.

Chan et al. also utilized a silicon-glass microfluidic device to monitor a cation exchange reaction in semiconducting CdSe nanocrystals by XANES at the Se K-edge (12.66 keV).156 In addition to silicon and glass, their device comprised a 2 μm thick silicon nitride/SU-8 X-ray window designed for performing measurements in fluorescence mode (Fig. 6a). In their experiment, a suspension of CdSe nanocrystals was introduced in a hydrodynamic flow focusing geometry surrounded by a sheath flow of Ag+ ions. As the ions diffused into the stream of nanocrystals, the kinetics of the CdSe → Ag2Se transformation could be followed over ∼100 ms with ∼8 ms time resolution owing to the narrow channel width, fast flow rates, and use of a microfocused X-ray beam. Similarly, Hofmann et al.157 and Tofighi et al.158 used a silicon-glass microfluidic device to study the synthesis of Au nanoparticles by fluorescence-based XANES at the Au L3-edge (11.92 keV). By utilizing on-chip turbulent cyclone mixers with <2 ms mixing time, the authors were able to gain access to early stages of the synthesis after only 1–2 ms of dead time.


image file: d4lc00637b-f6.tif
Fig. 6 Devices for hard XAS analysis. (a) Design of a microfluidic device with a silicon nitride window for XAS in fluorescence mode (adapted with permission from Chan et al., 2007; Copyright 2007 American Chemical Society).156 (b) Br K-edge XANES spectra of aqueous NaBr solutions at the indicated molar concentrations. The dark grey curves are from transmission detection and the green curves are from fluorescence detection (adapted with permission from Micheal Raj et al., 2021; Royal Society of Chemistry).159 (c) Design of a 3D-printed millifluidic device for XAS in transmission mode (adapted with permission from Dobrovolskaya et al., 2023; Copyright 2023 American Chemical Society).166

More recently, Micheal Raj et al. reported a silicon-glass type microfluidic device for performing both fluorescence and transmission mode XANES and EXAFS.159 They validated their device by studying Fe, Pb, and Br salt solutions and evaluating the quality of fluorescence vs. transmission mode data collected at different ionic concentrations between 1 mM and 1 M (Fig. 6b). Good quality data were obtained for Pb and Br solutions at the Pb L3-edge (13.04 keV) and Br K-edge (13.47 keV), however, the thick glass windows of the device (∼500 μm) resulted in strong attenuation at the Fe K-edge (7.11 keV), preventing further analysis. For Pb and Br, fluorescence detection performed better at concentrations of 1 mM, transmission better at 1 M, and both performed similarly at intermediate concentrations. The authors also paid particular attention to the world-to-chip connections of their device by designing a 3D-printed sample holder that enabled reproducible and safe mounting at the beamline. Britto et al. used a commercial microfluidic device for an operando XANES/EXAFS study of the synthesis of Pt NPs in fluorescence mode.160 Their device had a long channel length, and by tuning the flow rates, the authors were able to study the synthesis over reaction times from a few seconds to almost 40 minutes. Using multivariate analysis of the data, they were able to identify two intermediate phases that formed during the conversion of the H2PtCl6 precursor into metallic Pt.

Several polymer-based microfluidic devices have also been reported for XAS analysis at hard X-ray energies. For example, Zinoveva et al. utilized a PMMA microfluidic chip to study the synthesis of Co nanoparticles by fluorescence mode XANES at the Co K-edge (7.78 keV).161 Similarly, Oyanagi et al. studied the nucleation and growth of CdSe nanoparticles using fluorescence mode XANES and EXAFS at the Se K-edge.162 Their device consisted of two components: a microfluidic continuous flow mixer and a separate module comprising a Kapton capillary tube and a resistive heating element for studying reactions under high temperatures. Here, the combination of XANES and EXAFS enabled the modeling of the XANES data with multi-scattering calculations and comparison to EXAFS data in order to estimate the kinetics of Se–Cd bond formation. Monnier et al. developed a microfluidic electrochemical cell made from polyvinyl chloride (PVC) and brass for operando analysis of the reduction and oxidation of different iron-containing phases.163,164 They performed XANES and EXAFS at the Fe K-edge using both transmission and fluorescence read-out in order to study the corrosion of archeological samples and materials for the storage of nuclear waste.

4.2.2 Sensitivity considerations. One problem in early microfluidic XAS, and especially EXAFS, was low sensitivity to dilute species—even in fluorescence mode—because of the short beam path through the solution. This often results in total scan times nearing or exceeding one hour to obtain sufficient signal-to-noise ratio. Attempts have been made to solve this problem in different ways. For example, Krishna et al. utilized a commercial polyester (PET) millifluidic device with wider channels to obtain a higher fluorescence yield.165 The authors then employed this device to study the growth of nanostructured Au catalysts by XANES and EXAFS at the Au L3-edge. Alternatively, Dobrovolskaya et al. developed a 3D-printed millifluidic device for transmission-based XAS studies (Fig. 6c).166 An 80 mm pathlength was obtained by illuminating along straight sections of a serpentine flow channel rather than across the shortest dimension of the channel as is standard practice. Such long pathlengths combined with highly penetrating X-rays at the Pd K-edge (24.35 keV) allowed the authors to perform operando XANES and EXAFS of the formation of Pd nanoparticles with scan times of only 10 min at a second-generation synchrotron. However, averaging the signal along 80 mm of the flow channel significantly limited the potential time resolution of the device.
4.2.3 Soft X-rays. Researchers have also explored performing microfluidic XAS in the soft and tender X-ray regime, where transmission is lowered and increased interaction between the beam and device materials must be taken into account. To our knowledge, the first such example comes from Fulton et al., who modified the high-pressure cell of Hoffman et al.153 with thinner windows and a shorter beam pathlength.167 This enabled them to use the device down into the tender X-ray regime in transmission mode, but not to lower energies. Subsequently, Fuchs et al. reported a lower pressure flow cell made from PTFE.168 This cell was sealed by either a top silicon nitride (SixNy) or silicon carbide membrane designed for fluorescence read-out of XAS data in the soft X-ray regime. They also performed X-ray emission spectroscopy (XES), which is equivalent to XRF with higher energy resolution. Nagasaka et al. utilized a flow-cell comprising SixNy windows to perform transmission XANES measurements of liquid water at the O K-edge (532 eV).169 Performing such low energy XAS in transmission mode was made possible by keeping the fluid layer <1 μm by controlling the spacing of the silicon nitride membranes. Schreck et al. developed a similar flow-cell and used it to study not only pure water, but also solutions of CoCl2 and water–ethanol mixtures.170

More recently, Probst et al. utilized a PDMS-based droplet microfluidic device with a SixNy window to study the crystallization of CaCO3.171 They monitored the precipitation of amorphous calcium carbonation (ACC) over the first few seconds of the reaction by fluorescent XANES at the Ca K-edge (4.04 keV). However, owing to the background from the oil phase and the small droplet volumes, long scan times were required at each device position to obtain good photon counting statistics (>1 h). Additionally, the tender incident X-rays produced discolorations in the PDMS layer of the device, although the device shape and flow behavior were unaffected. Similarly, Brenker et al. utilized PDMS- and silicon-based droplet microfluidic devices with SixNy windows for fluorescence mode XANES at the Ca K-edge and found that the silicon devices were more resilient to the incident beam.172 The authors also used the Ca Kα line fluorescence yield to distinguish between droplets and the oil phase and isolate spectra from droplets. Despite this, the low total fluorescence yield still required the averaging of several long scans to obtain a good signal-to-noise ratio (>1 h), which demands highly stable device operation over long durations and large sample volumes.

4.2.4 Multi-technique analysis. Finally, a large focus of research has been in developing single devices that could support not only XAS analysis, but also accommodate other X-ray methods or simultaneous acquisition of non-X-ray data. In these cases, an extra layer of difficulty is in selecting a window material that works for different techniques or having multiple analysis windows made of different materials. For example, Chiarello et al. developed a millifluidic gas flow-cell to support simultaneous transmission mode XAS and diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) on catalyst powder beds (Fig. 7a).173 Since most IR transparent windows are made of crystalline materials that introduce diffraction artefacts in XAS spectra, the authors drilled a hole in their CaF2 window and infilled it with an amorphous carbon glue to facilitate transmission of the X-ray beam (Fig. 7b).
image file: d4lc00637b-f7.tif
Fig. 7 Hybrid devices for XAS and other techniques. (a) Exploded view of a millifluidic gas flow-cell for transmission mode XAS and simultaneous IR spectroscopy (adapted from Chiarello et al. 2014 with the permission of AIP Publishing).173 (b) EXAFS spectra of a Pt/Al2O3 catalyst obtained from the device in (a) with CaF2 windows (top) and CaF2 windows comprising a high temperature (HT) carbon glue bypass for X-ray transmission (bottom). Diffraction from the crystalline CaF2 window produces artefacts in the EXAFS data, which are eliminated by using the HT glue bypass (adapted from Chiarello et al. 2014 with the permission of AIP Publishing).173 (c) Conceptual design of an ultra-fast mixing device for XAS/SAXS/UV-vis analysis of nanoparticle synthesis. The inset shows the design of the butterfly mixing element (adapted with permission from Ramamoorthy et al., 2024; Royal Society of Chemistry).179

Conversely, Dann et al. designed an aluminum XAS/DRIFTS reactor with a separate CaF2 window for conducting IR analysis and used it to study a Pd–Al2O3 catalyst operating at high temperature.174 A polychromatic X-ray beam passed directly through a thinned section of the Al reactor, where it was used to perform energy dispersive XANES and EXAFS at the Pb K-edge and collect full spectra in only 4.8 ms (for XANES). The gas effluent at the outlet of their device was also directed towards an instrument for performing simultaneous mass spectrometry (MS). Similarly, Venezia et al. developed a silicon-glass millifluidic chip for XAS/DRIFTS/MS analysis that had separate thin silicon windows for the X-ray and IR beams, respectively.175 They also studied the operation of a Pd–Al2O3 catalyst and its performance in two separate reactions. Huyke et al. developed a microfluidic device for performing XAS and XES.176 Their device consisted of a fused silica hydrodynamic flow focusing mixer connected to a Kapton capillary for X-ray analysis. The authors used it to study the reduction of ferricyanide by ascorbic acid with millisecond time resolution.

Several devices have also been designed to support both XAS and SAXS analysis. For example, Karim et al. studied the synthesis of Pd nanoparticles in a silicon-glass microfluidic device with SAXS and fluorescence mode XANES/EXAFS at the Pd K-edge.177 For SAXS analysis, the glass layer was replaced by another Si layer to lower background scattering. Binninger et al. also performed XAS and SAXS with a millifluidic electrochemical flow cell.178 Their device was made from a polyether ether ketone (PEEK) housing that held two electrically conductive Kapton films to serve as both the X-ray windows and electrodes. Using this device, they performed Ir L3-edge (11.22 keV) transmission mode XANES/EXAFS to follow the oxidation state of a mixed Pt/IrO2–TiO2 electrocatalyst as a function of the applied electric potential. In turn, anomalous SAXS was used to follow the degradation of the Pt cathode, where performing SAXS at energies near the Pt L3-edge enabled isolation of the scattering from the Pt nanoparticles from that of the IrO2–TiO2 support. Recently, Ramamoorthy et al. demonstrated an ultra-fast microfluidic mixer for performing XAS, SAXS, and UV-vis analysis with dead times as low as 200 μs (Fig. 7c).179 They used the OSTEMER and Kapton-based device to study the nucleation and growth of Au NPs with transmission mode XANES at the Au L3-edge. With the aid of these three complementary techniques, the authors revealed the formation of transient pre-nucleation clusters and Au(I) lamellar phases prior to the nucleation of Au(0) nanoparticles. Notably, the authors performed both continuous and stopped-flow analysis on-chip to access shorter and longer reaction times, respectively.

4.3 XPCS

X-ray photon correlation spectroscopy (XPCS) has also received some attention in continuous flow micro- and milli-fluidic devices. XPCS is not a true spectroscopy technique based on the measurement of light transmission as a function of wavelength, but rather it is a time spectroscopy.187 Intensity fluctuations, or so-called “speckles”, appearing in X-ray scattering patterns collected over time using coherent X-ray radiation are temporally correlated to extract information on the physical dynamics of a sample, such as related to diffusive or advective motion.182 More simply, the technique is an extension of Dynamic Light Scattering (DLS) into the X-ray regime, where it is useful for analyzing thicker, more turbid samples at smaller spatial scales.188

The technique was first performed in flow by Fluerasu and co-workers, who used a simple millifluidic capillary cell to determine the diffusive dynamics of PMMA and latex nanoparticle suspensions.180,181 The authors found that scattering in the direction of the wavevector, q, perpendicular to the flow was insensitive to the advective component of particle motion at low shear rates (“transverse flow geometry”), whereas scattering in the direction parallel to the flow was strongly affected by advection (“longitudinal flow geometry”; Fig. 8). They thus exploited the transverse flow scattering to isolate the thermal motion of the particles and extract diffusion constants. Later, Fluerasu et al. followed up their work by using both the transverse and longitudinal scattering to study the diffusive and advective dynamics of a flow of silica nanoparticles.182 Transverse flow scattering data were used to deconvolute the longitudinal scattering data collected from different radial positions along the cross-section of the flow tube in order to calculate the shear relaxation rate. The rates calculated at each position matched well with a Poiseuille model of the flow.


image file: d4lc00637b-f8.tif
Fig. 8 An example of an XPCS flow experiment. Scattering of the incident beam (ki) in the longitudinal (q||) and transverse (q) directions by fluid flowing within the capillary is obtained from averaging of the regions within the dotted lines along the Z and X axes of the area detector, respectively. The photo shows the simple capillary setup of Fluerasu et al. (2008). Adapted with permission of the International Union of Crystallography.181

More recently, Urbani et al. further miniaturized this technology using a microfluidic chip and a microfocused X-ray beam.183 They mapped the advective dynamics at different channel positions along the chip, including straight and curved sections and a Venturi flow constriction. Flow velocities at the constriction calculated from XPCS data were compared to a CFD model with relatively good agreement, however some deviations were observed, especially at the center of the constriction where velocities were the highest. Lhermitte et al. improved the ability to calculate absolute flow velocities from XPCS data by performing heterodyne analysis, i.e., collecting scattering from the dynamic sample and a static reference material simultaneously.184 This heterodyne XPCS technique was subsequently utilized by Steinrück et al. to measure ion transport within the polymer electrolyte of a lithium-ion battery cell.185 Finally, Muhunthan et al. recently demonstrated a millifluidic flow-cell for performing split-pulse XPCS (sp-XPCS), X-ray pump X-ray probe (XPXP), and SAXS/WAXS measurements under extreme conditions at both synchrotrons and XFELs.186 The cell was constructed from titanium alloy with diamond windows and validated by obtaining speckle patterns from supercritical H2O at 380 °C and 25 MPa.

5 Devices for X-ray imaging

5.1 Brief theory and overview

The field of X-ray imaging comprises a wide variety of techniques that can provide morphological and chemical information on extended samples. These techniques can be divided broadly into two categories: full-field and scanning.189 Full-field techniques make use of an X-ray beam that irradiates an entire sample or region of interest in transmission mode. Conversely, with a scanning technique, a sample is irradiated by a small X-ray beam in a 2D raster pattern, and the image is constructed as a mosaic of the data collected at each position.189,190 Contrast in the image is obtained by recording the absorption or phase shift of X-rays by the sample using a 2D detector. Both of these quantities are related to a material's complex refractive index, n:
 
n = 1 − δ + iβ (3)
where δ is the real phase and β is the imaginary absorption of the sample material.191 Importantly, β is related to the linear absorption coefficient (μ) of the material, and thus, similar to XAS, X-ray imaging can be used to probe the chemical state of the sample by varying the wavelength of the incident beam (see eqn (2)).189 For obtaining contrast between soft materials or materials of similar densities, it is often necessary to probe the phase shift or utilize X-ray energies at an absorption edge of a target element.

With a scanning technique, data can be collected using different geometries and may be composed of single-wavelength absorption contrast measurements, or more commonly, of measurements collected at multiple wavelengths. Scanning instruments making use of other types of detection are also possible and can be used to construct, for example, fluorescence, scattering, or diffraction “maps”. Most techniques can also be made 3D by collecting 2D data (called “projections”) from multiple orientations of a sample under rotation and reconstructing the 3D volume using tomographic algorithms.192

In science and engineering applications, full-field X-ray imaging techniques relying on either phase-shift or absorption contrast are collectively referred to as full-field transmission X-ray microscopy (TXM), while scanning techniques are collectively referred to as scanning transmission X-ray microscopy (STXM). In both cases, the maximum spatial resolution is about 20–50 nm, mainly determined by the imaging zone plate in TXM, and by the focusing zone plate (which dictates the minimum raster step) in STXM. This resolution may be degraded in liquid though, in particular due to Brownian motion.

Owing to the wide variety of specific imaging techniques, this section will be organized by the type and application of the micro/millifluidic device rather than by the specific measurement technique as in the previous two sections. In fact, despite the number of beamlines and devices utilized, most studies have been conducted using similar device styles for similar applications. Specifically, the majority of papers report either thin microfluidic devices for performing TXM and STXM, often for electrochemistry applications, or larger millifluidic flow cells for performing tomographic studies of fluid flow and geochemistry within porous media. More details on all the papers reviewed below can be found in Table 3.

Table 3 Micro- and milli-fluidic devices for X-ray imaging
X-ray technique(s) Device type Device material(s) Fabrication and/or assembly method Window material, thickness Geometry, beam pathlength Sample(s) investigated Conditions Beamline, source, X-ray energy Beam size/resolution Acquisition mode and exposure time Ref.
STXM and XANES Microfluidic static cell Silicon Clamping Silicon nitride, 2 × 100 nm Transmission, ∼3 μm Clay Ambient temperature & pressure X1-A, NSLS, C K-edge (283.8 eV) Unknown Unknown Neuhausler et al. (2000)193
TXM Microfluidic static cell Polyimide, silicon Clamping Si/SixNy-coated polyimide, 2 × 150 nm Transmission, ∼10 μm CaCO3 Ambient temperature & pressure IRP, BESSY I, “water window” between 282–533 eV ∼40 nm resolution Multiframe, 2–10 s per exposure Rieger et al. (2000)194
Spectro-STXM and XANES Microfluidic electro-chemical cell Silicon, poly(chloro-trifluoroethyl-ene) Epoxy, vacuum grease Silicon nitride, 2 × 75 nm Transmission, ∼1 μm Polyaniline thin film Ambient temperature & pressure 5.3.2, ALS, C K-edge (283.8 eV) and N K-edge (401.6 eV) ∼50 nm resolution Unknown, <60 s Guay et al. (2005)195
Spectro-STXM Millifluidic gas flow cell Silicon, PDMS, glass Photolithography, wet etching, plasma etching, sputter coating, micromachining, plasma bonding, adhesive Silicon nitride, 2 × 100 nm Transmission, 0.8 mm Cu catalyst 260 °C, ambient pressure 11.0.2, ALS, Cu L3-edge (931.1 eV) ∼40–100 nm resolution Multiframe, ∼12 s per image per energy, full spectra in ∼15 min Drake et al. (2004)196
Spectro-STXM Microfluidic gas flow cell Silicon, aluminum, stainless steel, viton Machining, clamping, glue/wax Silicon nitride, 2 × 50 nm Transmission, ∼200 μm NaBr and diesel soot 5–27 °C, ambient pressure PolLux, SLS, O K-edge (532 eV) 30–40 nm beam spot Unknown Huthwelker et al. (2010)197
Ammonium sulfate Zelenay et al. (2011)198
Spectro-STXM Microfluidic gas flow cell Silicon, PEEK or aluminum, viton, PTFE Machining, clamping, epoxy Silicon nitride, 2 × unknown Transmission, ∼0.5–1.5 mm NaCl Ambient temperature & pressure 11.0.2 and 5.3.2.2, ALS, O K-edge (532 eV) 25 nm outer zone plate width Multiframe, ∼1 ms per pixel Kelly et al. (2013)199
Spectro-STXM Microfluidic gas flow cell Silicon, glass, brass, viton Machining, microfabrication, clamping Silicon nitride, 2 × 10 nm Transmission, ∼4 μm Fe-based FTS catalyst 25–350 °C, ambient pressure 11.0.2, ALS, C K-edge (283.8 eV), O K-edge (532 eV), and Fe L3- and L2-edges (708.1 and 721.1 eV) ∼40 nm resolution Unknown de Smit et al. (2008)200
Spectro-STXM Microfluidic gas flow cell Silicon, glass, brass, viton Machining, microfabrication, clamping Silicon nitride, 2 × 10 nm Transmission, ∼50 μm Fe-based FTS catalyst 25–500 °C, ambient pressure 10ID1, CLS, O K-edge (532 eV), and Fe L3- and L2-edges (708.1 and 721.1 eV) ∼40 nm resolution Unknown de Smit et al. (2009)201
Spectro-STXM Microfluidic gas/liquid flow cell Silicon, fluoropolymer O-ring Photolithography, reactive ion etching, chemical vapor deposition, wet etching, E-beam evaporation, clamping Silicon nitride, 2 × 50 nm Transmission, ∼500 nm CexTiO2-supported Pt catalyst 140 °C, ambient pressure 7.0.1.2, ALS, Ti L3- and L2-edges (455.5 and 461.5 eV), Ce M5-edge (883.8 eV), and Pt L3-edge (11.56 keV) ∼55 nm resolution Multiframe, 0.1–2 ms per pixel Yoo et al. (2020)202
TXM Microfluidic continuous flow cell Silicon, stainless steel, unknown O-ring material Machining, clamping, glue Silicon nitride, 2 × 100 nm Transmission, unknown Ag nanowires reacting with Au Ambient temperature & pressure Unknown beamline, APS, 12 keV ∼25 nm resolution Multiframe, 0.5 Hz Sun and Wang (2011)203
Spectro-STXM and μXAS Microfluidic electro-chemical cell Silicon, SU-8, NOA 84 Photolithography, UV curing Silicon nitride, 2 × Unknown Transmission, unknown Co-polypyrrole electrocatalyst Ambient temperature & pressure TwinMic, Elettra, Co L3-edge (778.6 eV) Unknown Unknown Bozzini et al. (2014)236
Spectro-STXM Microfluidic electro-chemical cell Silicon, fluoropolymer O-ring Photolithography, reactive ion etching, chemical vapor deposition, wet etching, E-beam evaporation, clamping Silicon nitride, 2 × 75 nm Transmission, ∼1 μm LiXFePO4 crystals Ambient temperature & pressure 11.0.2.2 and 5.3.2.1, ALS, Fe L3-edge (708.1 eV) 50 nm beam spot Multiframe, 1 ms per pixel Lim et al. (2016)204
Spectro-STXM Microfluidic electro-chemical cell Silicon, fluoropolymer O-ring Photolithography, reactive ion etching, chemical vapor deposition, wet etching, E-beam evaporation, clamping Silicon nitride, 2 × 100 nm Transmission, ∼1 μm Co(OH2) crystals Ambient temperature & pressure 11.0.2, ALS, Co L3-edge (778.6 eV) 50 nm resolution Multiframe, 2 ms per pixel + 2 ms delay, full spectra in ∼30 min Mefford et al. (2019)205
Mefford et al. (2021)206
Spectro-STXM Microfluidic electro-chemical cell and microfluidic gas flow cell Silicon, viton, unknown housing material Machining, clamping, tape Silicon nitride, 2 × 100 nm Transmission, ∼20 μm FeSO4 solution, polymer Ambient temperature & pressure BL4U, UVSOR-III, C K-edge (283.8) and Fe L3-edge (708.1 eV) ∼40–50 nm resolution Unknown Ohigashi et al. (2016)207
Spectro-STXM Microfluidic electro-chemical cell Silicon, acrylic polymer resin, unknown O-ring material 3D printing, proprietary techniques, clamping Silicon nitride, 2 × 50 nm Transmission, ∼1–1.5 μm Cu deposition on Au electrode Ambient temperature & pressure 10ID1, CLS, Cu L3-edge (931.1 eV) ∼25 nm resolution Multiframe, ∼20 s per image Prabu et al. (2018)208
Spectro-STXM Microfluidic continuous flow reactor Silicon, PDMS, PTFE, stainless steel Machining, soft lithography, clamping Silicon nitride, 2 × 50 nm Transmission, 3 μm CaCO3 with polymer additive Ambient temperature, ≤1 bar pressure drop PolLux, SLS, Ca L3- and L2-edges (346.4 and 350 eV) ∼43 nm resolution Multiframe, ∼5–50 ms per pixel, ∼5–11 min per image Gosse et al. (2020)89
HERMES, SOLEIL, 510 eV and Ca L3- and L2-edges (346.4 and 350 eV) ∼37–60 nm resolution
X-ray PIV (2D) Millifluidic continuous flow PTFE tubing Unknown PTFE, unknown Transmission, 750 μm Alumina microspheres Ambient temperature & pressure 1B2, Pohang Light Source, “white beam” 12.3 μm resolution Multiframe, 50 Hz Lee and Kim (2003)209
X-ray PIV (3D) Microfluidic continuous flow Silicone tubing Unknown Silicone, unknown Transmission, 490 μm Aluminum or solder microparticles Ambient temperature & pressure XOR-32-ID, APS, 18 keV 3.87 μm pixel size Multiframe, 60 Hz Lee et al. (2011)210
μCT Millifluidic flow cell Polyethylene, unknown Unknown Polyethylene, unknown Transmission, 6.5 mm Multiphase flow in porous polyethylene Ambient temperature & pressure X2B, NSLS, unknown 4.1 μm voxel size Multiframe, collection of tomograms after each injection Prodanović et al. (2006)211
μCT Millifluidic flow cell Silicone, PTFE, epoxy resin Epoxy, compression fittings Silicone and PTFE, unknown Transmission, 9 mm Carbonated salt solution in limestone core Ambient temperature, 0–3 MPa pressure ID19, ESRF, 40 keV 6 μm resolution Multiframe, collection of tomograms at six time points Noiriel et al. (2007)212
Fast-μCT Millifluidic flow cell with integrated pumps Polycarbonate, unknown Unknown Polycarbonate, unknown Transmission, 4 mm Multiphase flow in sandstone Ambient temperature, <10 kPa TOMCAT, SLS, 21.25 keV 3 μm voxel size Multiframe, 12 ms per projection, 16.8 s tomogram acquisition Berg et al. (2013)213
Fast-μCT Millifluidic flow cell with integrated pumps Unknown Unknown Unknown Transmission, 4 mm Multiphase flow in Robuglas Ambient temperature, <10 kPa TOMCAT, SLS, 36 keV 2.11 μm pixel size Multiframe, 40 ms per projection, continuous collection over 12 min Armstrong et al. (2014)214
Fast-μCT Millifluidic flow cell PEEK Compression fittings PEEK, unknown Transmission, 4.8 mm Multiphase flow in glass bead column Ambient temperature, 517.1 kPa pressure I22, diamond, unknown 3.25 μm resolution Multiframe, 6 s tomogram acquisition Hasan et al. (2020)215
Fast-μCT Millifluidic passive wicking cell PMMA with organic binders Powder-based 3D printing No window Transmission, 2 or 3.5 mm depending on the orientation Water in porous microbead column Ambient temperature & pressure ID19, ESRF, 40 keV “pink beam” 1.1 μm pixel size Multiframe, 0.5 s tomogram acquisition with 12 s read-out time Piovesan et al. (2020)216
Fast-μCT Millifluidic gravity-feed flow cell PMMA Compression fittings PMMA, unknown Transmission, 6 and 25 mm Multiphase flow in sandstone gravel column Ambient temperature & pressure I12, diamond, “white beam” between 50 and 150 keV or monochromatic beam of unknown energy between 53 and 150 keV 2.5 to 3.8 μm voxel size Multiframe, 90–200 μs per projection, 0.05 to 0.5 s tomogram acquisition Dobson et al. (2016)217
Tomographic X-ray PIV Millifluidic high-pressure flow cell (Hassler core holder) Viton, PEEK Compression fittings Viton, PEEK, unknown Transmission, 4 mm Multiphase flow containing silver-coated hollow glass tracer particles in sintered glass filter and limestone Ambient temperature, 2 MPa pressure TOMCAT, SLS, glassy carbon and borosilicate filtered “white beam” 2.75 μm voxel size Multiframe, 0.5 ms per projection, 0.25–0.5 s tomogram acquisition Bultreys et al. (2024)218
μCT Microfluidic continuous flow mixer (Kenics mixer) IP-S resin Two-photon stereolithography, UV curing, epoxy IP-S resin, unknown Transmission, 200 μm Mixing of two aqueous phases Ambient temperature & pressure P05, PETRA III, 11 keV P05: 1.14 μm voxel size P05: multiframe, 2 s per projection, ∼2.5 tomogram acquisition Knoška et al. (2020)219
TOMCAT, SLS, 18 keV TOMCAT: unknown TOMCAT: 80 ms per projection
μCT Millifluidic high pressure flow cell Epoxy resin, other unknown materials Compression fittings Epoxy resin, unknown Transmission, 9 mm Carbonated salt solution in limestone core Ambient temperature, 0.13 MPa back pressure ID19, ESRF, 40 keV 6 μm resolution Multiframe, collection of tomograms at six time points Noiriel et al. (2013)220
μCT Millifluidic flow cell Quartz tube Sintering, compression fittings Quartz, 1.4 mm wall Transmission, 1.6 mm BaSO4 in microporous quartz column Ambient temperature & pressure 13-ID-B, APS, 22 keV 1.24 μm voxel size Multiframe, collection of tomograms every 24 min Godinho et al. (2016)221
Fast-μCT Millifluidic high-pressure flow cell (Hassler core holder) Stainless steel, aluminum, steel wire, silicone, heat-shrink tubing and O-rings of unknown material Machining, compression fittings Aluminum, silicone, shrink wrap, unknown Transmission, 3 mm Salt solution in olivine rock 200 °C, 10 MPa fluid pressure and 15 MPa confining pressure 2-BM, APS, 65 keV “pink beam” 1.47 μm pixel size Multiframe, 10 ms exposure per projection, 20 s tomogram acquisition Fusseis et al. (2014)222
X-ray lamino-graphy Microfluidic packed-bed reactor Silicon, glass (Pyrex) Photolithography, wet etching, anodic bonding Glass, unknown, silicon, unknown Transmission, 30 μm CaCO3 Ambient temperature & pressure ID19, ESRF, 26 keV 0.7 μm resolution Unknown Morais et al. (2023)223
μCT Millifluidic passive counter-diffusion cell Tygon tubing, heat-shrink tubing Heat-shrinking Heat-shrink tubing, unknown Transmission, 3 mm BaSO4 in shale Ambient temperature & pressure I13-2, diamond, 27.6 keV “pink beam” 1.6 μm voxel size Multiframe, 100 ms exposure per projection, 80 s tomogram acquisition Godinho et al. (2019)224
μCT and XRD-CT Millifluidic passive counter-diffusion cell Glass tubes, His-3 BAG heat-shrink tubing, silicone tubing Heat-shrinking No window Transmission, 2.8 mm CaSO4·xH2O in CPG rod Ambient temperature & pressure μCT: I13-2, diamond, unknown μCT: 1.6 μm pixel size μCT: multiframe, 100 ms exposure per projection, 80 s tomogram acquisition Anduix-Canto et al. (2021)225
XRD-CT: ID11, ESRF, 40 keV XRD-CT: 50 μm × 50 μm beam size XRD-CT: single-shot, 1 s exposure per grid point
Scanning μXRF and XAS Microfluidic continuous flow mixer PDMS, silicon, O-rings and holder of unknown materials Soft lithography, machining, clamping, compression fittings Silicon nitride, 100 nm Fluorescence, 87.2 μm Water and pyridine Ambient temperature & pressure BL3U, UVSOR-III, N K-edge (401.6 eV) and O K-edge (532 eV) 30 μm × 30 μm, beam size, 30–61 μm step size Unknown Nagasaka et al. (2019)226
Scanning μXRF, SAXS/WAXS, and XRD Microfluidic continuous flow mixer (and others) PDMS, Kapton, 3D-printed resin, O-ring of unknown materials Soft lithography, plasma bonding, 3D printing, clamping, gluing, compression fittings Kapton, 8 μm, PDMS, 30 μm Fluorescence, >500 μm Iron oxide NPs Ambient temperature & pressure LUCIA, Soleil, Fe K-edge (7.11 keV) 3.5 μm × 3.5 μm beam size, 5 μm step size Single-shot, 300 ms exposure per grid point Chaussavoine et al. (2020)227
Scanning μXRF and XAS Microfluidic continuous flow cell Silicon, AF-32 Eco glass Photolithography, reactive ion etching, machining, plasma ashing, muffle urnace firing, epoxy, UV epoxy, compression fittings Glass, 30 μm Fluorescence, >30 μm Fe–As–S geochemical reactions Ambient temperature & pressure 4-BM, NSLS-II, 14 keV (for μXRF mapping) 5 μm × 5 μm beam size, 5–45 μm step size Single-shot, 50–200 ms exposure per grid point Chen and Kocar (2021)228
Scanning nano-XRF Microfluidic electro-chemical cell Glass, NOA, PET, PDMS Photolithography, wet etching, sputter coating, machining, UV epoxy Glass, ≤1 μm, PET, 12 μm Fluorescence, ∼20–50 μm Electrodeposition of Ag/AgCl Ambient temperature & pressure Carnaúba, Sirius, 9.7–13.7 keV 600 nm × 600 nm beam size, 5 μm step size Single-shot, 22.5 ms per grid point, 62 s for full map Neckel et al. (2021)229
XFEL-CDI Microfluidic continuous flow cell Silicon, Kapton Unknown Silicon nitride, 2 × 200 nm Transmission, 12 μm Ag NPs Ambient temperature, ambient to vacuum pressure BL2, SACLA, 4 keV 100 nm × 100 nm beam size Single-shot, 10 fs Matsumoto et al. (2022)230


5.2 Devices for STXM and TXM

5.2.1 Early work (static cells). Microfluidic STXM began in the early 1990s for the study of cells and subcellular structures,231,232 and these first studies influenced the design of most subsequent devices for both TXM and STXM. Both techniques are often performed at soft X-ray energies (<2 keV) at which beam pathlengths of more than a few microns through water result in significant attenuation. Therefore, microfluidic devices comprising ultrathin silicon nitride (SixNy) windows spaced just a few microns apart are ideally suited for performing in situ TXM and STXM experiments (Fig. 9a). While these types of devices are very thin, the window width is normally on the order of a millimeter, allowing large regions of the flow to be imaged. To our knowledge, microfluidic studies of hard condensed matter samples were not reported until a few years later in the year 2000. For example, Neuhausler et al. studied static samples of clay aggregates193 within a similar Si/SixNy wafer-based device and Rieger et al. studied the crystallization of CaCO3 from pre-mixed solutions using a device with Si/SiO2-coated polyimide windows.194 Subsequently, Guay et al. upgraded the basic silicon chip design by patterning Au electrodes over the SixNy windows to enable electrochemical measurements and time-lapse imaging of samples under electrical stimulus.195 Many similar static electrochemical cells for TXM/STXM were developed over the succeeding years and more information on these can be found in earlier reviews.233,234
image file: d4lc00637b-f9.tif
Fig. 9 Microfluidic devices for operando TXM/STXM. (a) A STXM setup with a MEMS gas nanoreactor for studying catalysis at high temperature (top). The inset shows a detailed illustration of the Pt electrode design and the SixNy windows (bottom left) and an example of spectro-STXM data from a heterogeneous catalyst particle (bottom right) (reprinted with permission from de Smit et al. 2008; Copyright 2008 Springer Nature).200 (b) An electrochemical STXM setup based on a Si/SixNy chip for studying Li-ion battery particles. The inset shows the side view of the device and the ∼1 μm spacing between the SixNy windows (from Lim et al., 2016; reprinted with permission from AAAS).204
5.2.2 Gas flow cells. The first flow reactor for in situ STXM was developed by Drake et al. to study catalysts under controlled gaseous environments at high temperatures.196 Their reactor had ultrathin SixNy windows coated with a patterned aluminum thin film to provide local resistive heating. The channels of the device were fabricated from glass and PDMS and had a total depth of 0.8 mm, a beam pathlength made possible by the lower attenuation coefficient of gases compared to liquids. The authors studied the oxidation and reduction of a Cu catalyst at temperatures up to 260 °C in CO/He atmospheres using spectro-STXM, i.e., spatially resolved XANES. By obtaining a series of STXM images at energies between 926 eV and 937 eV, where the Cu L3-edge is located, the authors could monitor the relative presence of Cu(I) and Cu(II) within the catalyst particles. Huthwelker et al. developed a gas-cell for spectro-STXM that they used to monitor aerosol particles and environmental contaminants under gases with different relative humidities.197 Notably, their device comprised two parts that enabled ease of mounting and rapid sample exchange: a back piece containing the heating/cooling system, gas connections, and standard mounting fixtures and a front clip holding the sample. The device was then utilized for subsequent environmental studies of ammonium sulfate aerosols.198 Kelly et al. designed a similar device comprising SixNy windows and a body machined from either aluminum or PEEK.199 However, they improved upon the design of Huthwelker et al. by enabling temperature and humidity measurements directly on-chip.

De Smit et al. adapted a micro-electromechanical systems (MEMS) cell designed for transmission electron microscopy (TEM)235 to perform the first spectro-STXM study of a working catalyst (Fig. 9a).200 They studied an iron-based catalyst during a Fischer–Tropsch synthesis (FTS), which is used to convert CO and H2 gas into hydrocarbon products. Imaging the catalyst at the C K-edge, O K-edge and Fe L2- and L3-edges enabled them to quantify the different iron phases formed and investigate their contribution to the FTS reaction. The same team followed up this work by imaging a single catalyst particle at different FTS reaction temperatures up to 500 °C.201 Yoo et al. also adapted a commercial cell designed for TEM to study the oxidation of CO gas by a TiO2-supported Pt catalyst.202 By using STXM at the Ce M-edge, they showed that doping of the TiO2 support with Ce encouraged the formation of highly efficient Pt single atoms at CeOx–TiO2 interfaces.

5.2.3 Liquid flow cells. Although originally developed for the field of biology,231,232 most subsequent liquid flow cells for TXM/STXM have been developed for applications in electrochemistry. Sun and Wang constructed a stainless steel flow cell with silicon nitride windows to study a galvanic replacement reaction between Ag nanowires and aqueous Au species by TXM.203 However, this device contained no active electrical components. To our knowledge, the first active electrochemical flow-cell was reported by Bozzini et al.236 Like previous designs for static electrochemical cells, theirs was also based on a Si/SixNy chip, however, better device sealing and more consistent spacing between SixNy windows were obtained through the use of a UV-curable resin instead of glue or vacuum grease. They used their device to study the electrodeposition of a Co–polypyrrole electrocatalyst using both STXM and scanning μXAS with fluorescence read-out to perform XANES at particular points of interest.

Lim et al. used a commercial electrochemical flow-cell to study single Li-ion battery particles under charging and discharging cycles (Fig. 9b).204 STXM enabled them to observe spatial heterogeneities in Li composition arising from non-uniform rates of lithiation and delithiation, which could affect battery performance and safety. Similarly, Mefford et al. used the same electrochemical flow cell to study compositional heterogeneities and the resulting catalytic heterogeneities within electrocatalyst single crystals.205,206 Other microfluidic electrochemical cells for TXM/STXM have been reported for studying electrode–electrolyte interactions,207 including the hybrid device of Prabu et al., which combined a Si/SixNy type chip with a 3D-printed holder for making world-to-chip connections.208

In contrast to most of the flow devices we have reviewed for scattering and spectroscopy, which were designed for rapid fluid mixing, most in situ TXM/STXM—and even TEM—studies have not been focused on flow management. This is likely because the reactions of interest were initiated by localized heating or an applied voltage rather than by the mixing of chemical reactants, and simply ensuring the replenishment of an electrolyte was sufficient for device operation. In this context, Gosse et al. presented a Si/SixNy type device for STXM designed to study precipitation reactions with precise flow control.89 By using a pressure-actuated, rather than syringe-driven, system with flow meters upstream and downstream of the chip, the authors demonstrated the ability to adjust or stop flows within seconds. They also showed that the liquid within the Si/SixNy cell could be completely refreshed in less than two minutes using flows of only a few microliters per minute.

5.3 Tomography for flow imaging

5.3.1 Early two-dimensional and tomographic work. Two-dimensional X-ray imaging of macroscale flows dates back to the 1950s and 1960s, when it was used to visualize opaque multiphasic fluidized beds and chemical reactors that could not be imaged by optical methods.237 Much later work at third generation synchrotron sources on X-ray phase contrast imaging paved the way for higher resolution studies of weakly absorbing liquids and soft materials.238 Lee and Kim exploited these developments to perform the first 2D X-ray particle image velocimetry (PIV) study of a microflow in 2003.209 Using an X-ray beam, rather than a conventional laser-based PIV probe, the authors irradiated a flow of alumina microspheres within a 750 μm diameter PTFE tube to extract the velocity field. Lee et al. subsequently extended the technique into 3D by using a beam splitter and mirror to image a microflow from two directions and extract a third velocity component.210

Around the same time, work was beginning in the geological community to visualize processes within porous rocks that were also inaccessible to light microscopy. X-ray micro-computed tomography (μCT) was already an established method for investigating the internal 3D microstructure of geological samples. The next step was to combine it with a fluidic sample environment to perform dynamic studies. Cylindrical rock cores with diameters on the millimeter scale are thin enough to allow the transmission of hard X-rays, and they can also be assembled into simple millifluidic devices by sealing the sides with an epoxy resin and making fluid connections at both ends. Early work in this area required the sample to be alternatively applied with a flow and then taken to the beamline for imaging due to the difficulty of attaching flow equipment to a rotating tomography stage.239 Subsequent studies, such as from Prodanović et al.211 and Noiriel et al.,212 were performed in situ but not operando, in that the devices were mounted on the stage during flow, but the flows were stopped during imaging. Thus, the difficultly of connecting the flow apparatus to the sample stage and the long acquisition times required to obtain a full tomogram continued to limit the potential time-resolution of the technique.

5.3.2 Investigations of flow phenomena. The situation improved with the development of so-called “fast-μCT” methods, in which full tomograms can be acquired in seconds or less. This is made possible using synchrotron radiation and high frame rate detectors by synchronizing the rotation of the sample stage with the acquisition of 2D projections.240,241 Berg et al. were the first to utilize these advances in fast-μCT for operando studies of fluid transport within opaque porous media.213 To enable the fast rotation of the sample for the rapid acquisition of projections, the authors constructed an integrated millifluidic device containing pumps, fluid reservoirs, and an encased rock sample that could be mounted directly on the motorized stage without external tubing connections (Fig. 10a). They investigated the displacement of oil within a sub-millimeter porous network of sandstone and observed Haines jumps, snap-off events, and the entrapment of oil droplets with a temporal resolution of 16.8 s in 3D (Fig. 10b). Armstrong et al. later followed up this work to obtain better spatial and temporal resolution by incorporating information from individual 2D projections on the millisecond time-scale.214 The authors utilized the continuous rotation of the sample and simultaneous collection of projections to identify time points at which fluid motion occurred. Then the projections from intervals of relatively little motion could be reconstructed into 3D tomograms with fewer blurring artefacts.
image file: d4lc00637b-f10.tif
Fig. 10 Operando X-ray micro-computed tomography. (a) An integrated millifluidic device for fast-μCT that can be mounted on a rotated tomography stage without external connections (adapted with permission from Armstrong et al. 2014; Copyright 2014 John Wiley and Sons).214 (b) Drainage of water (gray) and infilling of oil (red) within sandstone pores observed with fast-μCT (1–2). (3) Cross-section at a pore throat showing the three phases (water, oil, and quartz rock) (used with permission from Berg et al. 2013; Copyright 2013 The Authors).213

The time resolution of full tomograms has continued to increase. For example, Hasan et al. used fast-μCT to study solute transport in water-saturated and -unsaturated porous media with a full-tomogram time resolution as short as 6 s.215 Piovesan et al. fabricated a porous millifluidic device using a powder-based 3D printer and used it to study capillary wicking in 3D.216 They performed fast-μCT with sub-second time resolution, however, the readout time of the detector limited the frequency of tomogram acquisition, requiring a 12 s time step between consecutive tomograms. To the best of our knowledge, Dobson et al. performed the first operando fast-μCT experiment with both a sub-second time resolution and time step, with full-tomogram acquisition frequencies up to 20 Hz.217 More recently, Bultreys et al. combined earlier work in 2D X-ray PIV with fast-μCT to perform the first synchrotron-based 3D X-ray PIV study of flow within porous media.218 They imaged multiphase flows containing tracer particles within limestone and sintered glasses and obtained tomograms with 0.25 s time resolution at an acquisition rate of 4 Hz. Many additional studies of hydrology in porous media have been conducted using fast-μCT and millifluidics and cannot all be covered here.242–247 To our knowledge, the only microfluidic μCT study of fluid transport was reported by Knoška et al., who characterized the flow within a helical Kenics mixer by merging streams of low-contrast water and a high-contrast KI solution.219 Due to the steady-state concentration profile of the flows in the static mixer, standard μCT with an acquisition time of ∼2.5 h was sufficient to capture the mixing process.

5.3.3 Investigations of geochemistry and mineral precipitation. In parallel with tomographic work focused on fluid transport, operando tomography was also used to investigate geochemical processes occurring at the fluid-pore interface. Unlike for fluid dynamics, due to the slow speed of many of these processes, it is not always necessary or practical to obtain consecutive tomograms of near-second resolution over durations of hours. This would result in the collection of large volumes of redundant data. For this reason, slower acquisition rates are often used, and collections may be spaced further apart throughout the experiment. For example, Noiriel et al. used μCT and a millifluidic device to study the dissolution of calcite within fractured limestone under acidic conditions. They followed the dissolution process over 55 hours using tomograms collected at six different time points.220 Similarly, Godinho et al. used a continuous flow millifluidic device to study the precipitation of barite (BaSO4) within a microporous silica column.221 Using μCT with a time resolution of 24 minutes, they followed the precipitation-induced occlusion of pores and the resulting changes in the flow velocity and crystal growth rate. Fusseis et al. developed a Hassler-type millifluidic cell for performing fast-μCT of fluid–rock interactions under extreme conditions up to 200 °C and 15 MPa.222 In particular, they studied the dissolution and precipitation of magnesium-based minerals under high salt and carbonate solution conditions. More recently, Morais et al. further miniaturized these geochemical studies by using a packed-bed microfluidic reactor and X-ray laminography,223 a computed tomography technique for obtaining high-resolution 3D images of thin planar samples.248 The authors packed a wide microfluidic reservoir with particles of calcite and subsequently injected packets of acidified media to study the dissolution dynamics of this mineral bed. X-ray tomograms were acquired between each injection of an acidified fluid volume, and the changing structure of the bed caused by the dissolution process was input into a CFD model to simulate the resulting flow profiles.

To investigate diffusive- rather than advective-driven transport processes in samples with smaller pore sizes, several researchers fabricated simple counter-diffusion-based devices. These devices are easier to set up at tomography beamlines than their flow-based counterparts since it is not necessary to mount pumps on the sample stage or connect devices to external equipment. For example, Godinho et al. made a passive-diffusion cell by connecting short sections of tubing to both sides of a 3 mm-diameter shale sample containing micrometer-scale fractures and sub-micron sized pores.224 The tubing sections served as fluid reservoirs containing counter-ions, which were allowed to diffuse into the sample, and the whole assembly was mounted vertically on the stage using one of the tubes as the support. Using this device, the authors continued their μCT studies of BaSO4. They observed that as Ba2+ and SO42− ions diffused and reacted within the sample, barite precipitated first within the larger fractures and then later in smaller fractures and pores.

Anduix-Canto et al. used a similar setup to study the precipitation of calcium sulfate within controlled porous glass (CPG) rods with an average pore diameter of 7 nm.225 The authors performed not only μCT, but also X-ray diffraction computed tomography (XRD-CT) to reveal both the morphology and crystal structure of the phases that precipitated over time. Utilization of CPGs enabled the authors to confine the solution within nanopores that were completely isolated from macropores, and this nanoscale confinement effect led to the precipitation and stabilization of normally unstable amorphous and hemihydrate CaSO4·xH2O phases. The field of geological μCT is a large area, often on the border between milli- and “macro”-fluidics, and thus cannot be completely covered here. The reader is directed to other reviews for further information.246,249

5.4 Other scanning techniques

In addition to absorption- and phase-contrast X-ray imaging and tomography, several microfluidic studies have made use of the spatially resolved acquisition of other types of X-ray data. The most popular approach in the physical sciences has been using scanning μXRF to map micromixers and visualize the concentration profiles of chemical species. Such an approach was implemented by Nagasaka et al., who utilized a microfluidic device comprising a PDMS microchannel and a silicon nitride window to study the mixing of pyridine and water.226 They first mapped the steady-state laminar microflow by μXRF and then obtained N K-edge XAS spectra at different points across the width of the microchannel to determine the local pyridine concentration. Similarly, Chaussavoine et al. developed a vacuum-compatible microfluidic device for studying the synthesis of iron (hydr)oxide nanoparticles.227 They performed μXRF of the channel and identified regions of interest at which to acquire full Fe K-edge XANES spectra using fluorescence read-out. This approach was also utilized in several studies previously reviewed in section 4.156,171,172

Scanning μXRF has also been used to study spatial heterogeneities within planar microfluidic devices. For example, Chen and Kocar investigated Fe–As–S geochemical reactions using devices containing arrays of quartz micro-posts as models of porous rock.228 First, the authors would mix a basal salts solution (BSS) with an Fe2+ solution to precipitate iron (hydr)oxides. Subsequently, As- and S-containing solutions were introduced along with NaBr as a flow tracer, and the resulting flow profiles and sorption of As/S were visualized by μXRF (Fig. 11). Similarly, Neckel et al. used scanning XRF to map the electrodeposition of Ag/AgCl films within a microfluidic electrochemical cell.229 They utilized a nanobeam to obtain maps of Ag nucleation sites with sub-micron resolution.


image file: d4lc00637b-f11.tif
Fig. 11 Scanning μXRF within a microfluidic device containing micro-post arrays. Brightfield optical image illustrating the flow direction and chemical addition (top). Model geochemical reactions in the Fe–As–S system are monitored using the appropriate Kα fluorescence peak (bottom and insets). Br is utilized as a flow tracer (reproduced with permission from Chen and Kocar, 2021; International Union of Crystallography).228

Finally, in an interesting recent study, Matsumoto et al. utilized a multi-window Si/SiyNx microfluidic device at an XFEL to perform coherent diffraction imaging (CDI) of nanoparticle aggregation.229 Two separate suspensions of spherical and rod-shaped gold nanoparticles, respectively, were stored on-chip. Mixing of the solutions was initiated by an XFEL pulse breaking a silicon nitride window at the end of the flow channel and rapidly depressurizing the device under vacuum. Once a flow was established, the device was scanned upstream along the channel to image particles under different aggregation states. To our knowledge, this may be the only example of on-chip microfluidic analysis at an XFEL, and it illustrates a creative way to overcome the challenge of X-ray pulse-induced device damage through experimental design.

6 Developments in laboratory-based analysis

In this final review section, we will look at developments in performing in situ micro- and milli-fluidics experiments in the ‘home’ laboratory, that is, using laboratory X-ray instruments rather than large-scale facilities. Even researchers and institutions with strong official links to a synchrotron radiation facility cannot expect more than a few weeks of beamtime per year under normal circumstances. For other researchers and relative newcomers to user facilities, more than a few days per year might be a luxury. Conducting an experiment with a complex sample environment at another institution also requires a great deal of planning, set-up, and hard work. Additionally, one may utilize unfamiliar equipment or software and rely heavily on the help of local staff, who will have varying levels of investment in user experiments. These are just a few of the many reasons why researchers are interested in the ability to perform microfluidics experiments—and in situ experiments in general250,251—in a laboratory setting.

Fortunately, there have been substantial improvements in laboratory X-ray sources and hardware over the past decade and a half. First and foremost, the flux of rotating anode and liquid metal jet X-ray sources has approached that of second-generation synchrotron facilities,11 achieving up to ∼109 photons per second at the sample depending on the source type and collimation/focusing optics.15,252 State-of-the-art sealed tube sources are also improving and able to reach the ∼107–108 range. Multilayer optics and scatterless slits have provided better quality beams,253,254 and sensitive, low-noise, hybrid photon counting detectors used at synchrotrons are now commonly employed on laboratory systems.255,256 Further, current commercial systems offer better software and increased functionality, automation, and sample control, reducing the need to build custom platforms. All these factors have increased the performance and user-friendliness of laboratory X-ray instruments, resulting in renewed interest in using them to perform operando experiments. While they still cannot compete with third- and fourth-generation synchrotrons in terms of flux or coherence, more and more micro- and milli-fluidics experiments are becoming feasible in the laboratory, as the papers reviewed below demonstrate (Table 4). Here, we will again cover progress in scattering/diffraction, spectroscopy, and imaging, and we will also include biological and soft matter applications since, to our knowledge, laboratory X-ray analysis in these areas has not been reviewed previously. Where possible we will also compare the data quality and time resolution of laboratory-based studies to similar synchrotron experiments.

Table 4 Micro- and milli-fluidic devices for laboratory-based X-ray analysis
X-ray technique(s) Device type Device material(s) Fabrication and/or assembly method Window material, thickness Geometry/mode, beam pathlength Sample(s) investigated Conditions Instrument, source, detector Beam size/resolution Acquisition mode, exposure time Ref.
SAXS Millifluidic continuous flow reactor Teflon tubing, quartz capillary, commercial micromixer Components connected by tubing, compression fittings Quartz capillary, 10 μm wall Transmission, 1–3.8 mm Au NPs and Ag NPs Ambient temperature & pressure SAXSess (Anton Paar), sealed tube source (PANalytical, 8.04 keV), CCD detector (Roper Scientific) Unknown Multiframe, 20 × 10 s Polte et al. (2010)257
Polte et al. (2012)258
SAXS Millifluidic continuous flow reactor Teflon tubing, quartz capillary, commercial micromixer Components connected by tubing, compression fittings Quartz capillary, 10 μm wall Transmission, 1 mm Pd NPs Ambient temperature & pressure SAXSess (Anton Paar), sealed tube source (PANalytical, 8.04 keV), CCD detector (Roper Scientific) Unknown Multiframe, unknown Kettemann et al. (2015)259
SAXS/WAXS Millifluidic continuous flow reactor Teflon, polyethylene tubing, quartz capillary Components connected by tubing, compression fittings Quartz capillary, 10 μm wall Transmission, 1 mm Au NPs 22, 33, and 45 °C, pressure Double Ganesha AIR (SAXSLAB), rotating anode source (Rigaku, 8.04 keV), hybrid-photon counting detector (Dectris) Unknown SAXS: multiframe, 60, 300, or 600 s exposures depending on the condition and reaction time Chen et al. (2015)260
WAXS: unknown
SAXS/WAXS Millifluidic continuous flow reactor Teflon, polyethylene tubing, quartz capillary Components connected by tubing, compression fittings Quartz capillary, 10 μm wall Transmission, 1 mm ZnO NPs 40 and 50 °C, pressure Double Ganesha AIR (SAXSLAB), rotating anode source (Rigaku, 8.04 keV), hybrid-photon counting detector (Dectris) Unknown SAXS: multiframe, 30, 60, or 600 s exposures depending on the reaction time Herbst et al. (2019)261
WAXS: multiframe, 1200 s exposures
SAXS Millifluidic electrochemical flow cell PEEK, Kapton (gold-coated) Machined, clamped Kapton, 2 × 50 μm Transmission, 50–123 μm of catalyst, 2 mm of electrolyte Pt NPs on a carbon support Ambient temperature & pressure X'Pert Pro (PANalytical), sealed tube source (PANalytical, 8.04 keV), 1D solid-state detector (PIXcel) Unknown Multiframe 7 × 1 h Tillier et al. (2016)262
SAXS Millifluidic high performance liquid chromoto-graphy system Unknown Compression fittings connecting multiple commercial devices Glass capillary, unknown wall Transmission, unknown Proteins Ambient temperature & pressure BioXolver L (Xenocs), liquid metal jet source (Excillum, 9.25 keV) ∼1 mm × 1 mm Single-shot, 60 s Bucciarelli et al. (2018)263
SAXS Millifluidic continuous flow mixer PMMA, quartz capillary Unknown Quartz capillary, unknown wall Transmission, 2 mm SiO2 NPs and human serum albumin Ambient temperature & pressure NanoStar (Bruker), microfocus source (8.04 keV), 2D gas-based detector (VÅNTEC-2000) ∼300 μm diameter Single-shot, 2.5–30 h Anaraki et al. (2020)264
SAXS Millifluidic continuous flow reactor Glass capillary, PTFE tubing, polymer fittings Compression fittings, tubing coiling Quartz capillary, 10 μm wall Transmission, 1 mm Iron oxide NPs 60 °C, ambient pressure SAXSess (Anton Paar), sealed tube source (GE, 8.04 keV), 1D solid-state detector (Dectris) 17 mm × 0.25 mm Single-shot, 60 s Besenhard et al. (2020)127
SAXS Microfluidic segmented flow OSTEMER 322 Photolithography, soft lithography, cured and laminated OSTEMER 322, 2 × 200 μm Transmission, 150 μm Device materials Ambient temperature & pressure Xeuss 2.0 (Xenocs), sealed tube source (Xenocs, 8.04 keV), hybrid photon counting detector (Dectris) Unknown Multiframe, 500 s total scan time Lange et al. (2020)92
Micro-XRD Microfluidic electro-chemical cell PVC, brass, graphite Machining, clamping, pressure fittings, conductive epoxy Kapton, unknown, graphite, 500 μm Transmission, ∼400 μm Iron and iron oxide phases Ambient temperature & pressure Custom diffractometer, rotating anode source (Rigaku, 17.4 keV), image plate detector (unknown) 400 μm diameter or 20 × 20 μm Multiframe, 20 min scan time Monnier et al. (2008)163
Monnier et al. (2014)164
SAXS Microfluidic evaporative flow PDMS Photolithography, soft lithography, plasma bonding PDMS, ∼150 μm Transmission, 25 μm Au colloidal supercrystals Ambient temperature & pressure Rotating anode source (Rigaku, 8.04 keV), hybrid-photon counting detector (Dectris) ∼0.5 mm × 0.5 mm Single-shot, ∼3.5 min García-Lojo et al. (2021)265
SAXS Millifluidic flow dialysis device COC, commercial dialysis insert 3D printing COC, 2 × 50 μm Transmission, 12 mm Lipids and polymers Ambient temperature & pressure Microfocus source (Xenocs, 17.4 keV), unknown detector Unknown Multiframe, 360 × 1 min Ehm et al. (2022)266
Single-shot, 20 min
SAXS/WAXS and PXRD Microfluidic segmented flow PMMA, PTFE, Kapton, silicone UV laser cutting, clamped, compression fittings Kapton, 2 × 75 μm Transmission, 300 μm CaCO3 with nucleating agents, silica NPs Ambient temperature & pressure Xeuss 2.0 (Xenocs), liquid metal jet source (Excillum, 9.25 keV), hybrid photon counting detector (Dectris) ∼250 μm × 250 μm Multiframe, 60 × 0.5 s Levenstein et al. (2022)15
Milli-fluidic segmented flow Kapton capillary Interference fit, compression fittings Kapton capillary, 100 μm wall Transmission, 1 mm Calcite and paracetamol crystals XtaLab Synergy R or Synergy Custom (Rigaku), microfocus rotating anode source (Rigaku, 8.04 keV), hybrid photon counting detector (Rigaku) ∼140 μm × 140 μm Multiframe, 3600 × 25 ms
SAXS Microfluidic continuous flow OSTEMER, Kapton Photolithography, PDMS injection molding, cured, cure-bonded OSTEMER, 2 × 50 μm and Kapton, 25 μm Transmission, 400 μm Au NPs, silica NPs, BSA protein, latex NPs Ambient temperature & pressure Xeuss 2.0 (Xenocs), microfocus source (Xenocs, 8.04 keV), hybrid photon counting detector (Dectris) 250 μm × 250 μm Single-shot, 10–60 min Radajewski et al. (2023)267
SAXS Millifluidic continuous flow reactor PFA tubing, glass capillary Compression fittings, tubing coiling Glass capillary, unknown wall Transmission, unknown Polymer nano-objects 75 °C, 6.9 bar backpressure Xeuss 3.0 (Xenocs), liquid metal jet source (Excillum, 9.25 keV), hybrid photo counting detector (Dectris) 0.4 mm diameter Single-shot, 5 min Guild et al. (2023)268
PXRD Various micro- and milli-fluidic devices Various Various Various Various Calcite, theophylline, KNO3, Na2SO4 Various XtaLab Synergy Custom (Rigaku), microfocus rotating anode source (Rigaku, 8.04 keV), hybrid photon counting detector (Rigaku) 150 μm diameter, tunable Multiframe, 18 s exposure per frame Turner et al. (2024)269
XANES Millifluidic photocatalysis cell Unknown body material, scotch tape 3D printing (DLP), clamping, tape Scotch tape, unknown Fluorescence, unknown Pt catalyst on TiO2 support Ambient temperature, Ar atmosphere R-XAS Looper (Rigaku), unknown source at Pt L3-edge, silicon drift detector Unknown Single-shot, ∼1–2 min Kozyr et al. (2023)270
XRF Capillary electro-phoresis cell Fused silica capillary with polyimide coating Capillary mounted in plastic housing Fused silica capillary with polyimide coating, 33.5 μm wall thickness Fluorescence, unknown Free and complexed metal ions Ambient temperature & pressure Eagle II (EDAX), Rh target excitation source (EDAX), SiLi detector (EDAX) ∼50 μm spot Multiframe, 10 s dwell time for full spectra Miller et al. (2003)271
μXRF and XRD Miniaturized XRF chip Glass, PDMS, lead Various Glass, unknown Fluorescence, unknown Metal foils Ambient temperature & pressure XRD spectrometer with HPGe detector (Canberra Industries) Uncollimated radiation from Am source Unknown Greaves and Manz (2005)18
Various millifluidic XRD chips Polycarbonate, PDMS, borosilicate glass, Various materials of unknown thickness Transmission, unknown Corundum powder D8 Discover (Bruker), sealed tube (8.04 keV), 2D GADDS detector 1 mm diameter
XRF Microfluidic liquid–liquid extraction device Kapton, PLA 3D printing, epoxy, glue Kapton, 7 μm Fluorescence, >200 μm La, Eu, and Yb ions in aqueous and organic media 20–35 °C, 60 mbar back pressure in aqueous channel X-ray spectrometer with sealed tube source (Moxtek, 22.16 keV), X-123 SDD detector (Amptek) Unknown Multiframe, 120 s per exposure Maurice et al. (2022)272
XRF Millifluidic solid-phase extraction device Kapton tubes, PEEK tubing, 3D-printed holder 3D printing, tape, silicone sealant Kapton tubes, 38 μm wall thickness Fluorescence, 1.52 mm La, Nd, Yb and Fe ions in acidic media Ambient temperature & pressure X-ray spectrometer with sealed tube source (Moxtek, 22.16 keV), X-123 SDD detector (Amptek) Unknown Multiframe, 123 s per frame Olivier et al. (2023)273
μCT Millifluidic flow cell (Hassler core holder) PEEK body with other unknown materials Compression fittings PEEK, 2 mm Transmission, 6 mm Multiphase flow in sandstone Ambient temperature & pressure Unknown 5 μm voxel size Multiframe, 80 min tomogram acquisition Youssef et al. (2009)274
μCT Millifluidic high-pressure flow cell (Hassler core holder) M55 carbon fiber, viton, aluminum, Kapton Compression fittings Carbon fiber, viton, aluminum, and Kapton, unknown thickness Transmission, 4–6 mm Supercritical CO2 and brine in limestone 50–63 °C, 10 MPa pore pressure, 11–13.1 MPa confining pressure Versa XRM-500 (Xradia), unknown source and detector 2–6.4 μm voxel size Multiframe, continuous collection of projections, 75–90 min tomogram acquisition Andrew et al. (2013)275
Andrew et al. (2014)276
Andrew et al. (2014)277
Fast-μCT Millifluidic flow cell (Hassler core holder) PMMA, viton, viton O-rings, hydrophobic membranes of unknown material Compression fittings PMMA, viton, unknown thickness Transmission, 6 mm Multiphase flow in sandstone Ambient temperature, unknown pressure Custom gantry-based scanner, microfocus sealed tube source, 2D CMOS scintillator detector 7.4–14.8 μm voxel size Multiframe, continuous collection of projections, 12 s or 21 min tomogram acquisition Bultreys et al. (2016)278
μCT Millifluidic high-pressure flow cell (Hassler core holder) M55 carbon fiber, steel, viton, aluminum, Kapton Compression fittings Carbon fiber, viton, aluminum, and Kapton, unknown thickness Transmission, 4 mm Multiphase flow in limestone 50 °C, 10 MPa pore pressure, 13 MPa confining pressure Versa XRM-500 (Xradia), unknown source and detector 3.8 μm resolution Multiframe, ∼15 min tomogram acquisition Menke et al. (2015)279
μCT Millifluidic flow cell Glass, heat shrink tubing Sintering, compression fittings, heat shrink Heat-shrink tubing, unknown Transmission, 6 mm BaSO4 in microporous glass column Ambient temperature & pressure Custom scanner, unknown source, 2D detector (PerkinElmer) 4.3 μm voxel size Multiframe, 0.25, 1.5 and 4.9 h tomogram acquisition depending on number of projections selected Gajjar et al. (2018)280
Helical-CT Millifluidic flow cell Quartz Sintering, compression fittings No window Transmission, 3.9 mm CaCO3 in microporous glass column Ambient temperature & pressure Heliscan (FEI), unknown source and detector 2.24 μm voxel size Multiframe, ∼2.5 h tomogram acquisition Godinho et al. (2018)281
μCT Millifluidic packed-bed cell PEEK Compression fittings, grain packing PEEK, unknown Transmission, 7.5 mm CaCO3 grains Ambient temperature & pressure NSI XCT scanner (North Star Imaging), unknown source, 2D detector (PerkinElmer) 19.3 μm voxel size Multiframe, 83 s tomogram acquisition Singh et al. (2024)282
Tomographic X-ray PIV Millifluidic flow cell and microporous column Carbon fiber tube Sintering, compression fittings Carbon fiber, 1 mm Transmission, 6.35 mm Silver-coated hollow glass spheres in different creeping flows Ambient temperature & pressure CoreTOM (TESCAN), sealed tube source, 2D silicon detector 18 μm voxel size Multiframe, 14.5 ms per projection, <3 s tomogram acquisition Mäkiharju et al. (2021)283
Tomographic X-ray PIV Millifluidic flow cell (Hassler core holder) Viton and other unknown materials Sintering, compression fittings, grain packing Unknown Transmission, 4 mm Silver-coated hollow glass spheres in a sand pack and microporous glass column Ambient temperature, 2 MPa confining pressure Custom gantry-based scanner, microfocus sealed tube source, 2D CMOS scintillator detector 11.8 μm voxel size Multiframe, 100 ms per projection, 70 s tomogram acquisition Bultreys et al. (2022)284
GE-XRF and scanning μXRF Passive semi-open microfluidic chip Unknown Unknown fabrication steps, acidic surface treatment, silanization No window Fluorescence, dried sample thickness Dried Cu, Cd, and Fe salts Ambient temperature & pressure Sealed tube source (17.4 keV), silicon drift detector (Rontec) GE-XRF: ∼10 mm GE-XRF: single-shot, 300 s Tsuji et al. (2005)285
μXRF: 50 μm spot μXRF: unknown
Confocal 3D-XRF Multi-layer microfluidic chip PET Commercially purchased PET, 2 × 125 μm Fluorescence, 849 μm Aqueous Co and Cu solutions Ambient temperature & pressure Ceramic sealed tube source (RTW, 17.4 keV), silicon drift detector (Bruker) 30–40 μm spot, 30 μm step size Single-shot, 50 s per grid point Nakano and Tsuji (2010)286
Scanning μXRF Passive microfluidic chip Kapton, polycarbonate, silicone transfer tape CO2 laser cutting, tape Kapton, 40 μm Fluorescence, ≥1 mm Sr solutions Ambient temperature & pressure Custom μXRF system, sealed tube source (XOS, 20.22 keV), silicon drift detector (Hitachi) Custom: 200 μm diameter, 200 μm step size Custom: single-shot, 1 s per grid point McIntosh et al. (2014)287
Eagle III (EDAX): sealed tube source (20.22 keV), Si(Li) detector Commercial: 50 μm spot, 85 μm step size Commercial: single-shot, 0.2 s per grid point


6.1 X-ray scattering and diffraction

6.1.1 Early work and acquisition time considerations. Micro- and milli-fluidic laboratory-based X-ray scattering analysis has its origin in the recirculating flow cells used to sample materials forming within batch reactors. This early work focused on monitoring the crystallization of model pharmaceutical compounds by powder X-ray diffraction288,289 or the synthesis of nanoparticles by SAXS,290 where the continuous recirculation of products between mL- to L-scale batch reactors and the X-ray instrument facilitated operando measurements. To our knowledge, it was Polte et al. who first coupled a true continuous flow reactor to a laboratory X-ray instrument for operando analysis (Fig. 12a).257 The authors utilized a millifluidic mixer module connected to a glass capillary to study the synthesis of gold NPs by SAXS. By using different lengths of tubing between the mixer and analysis capillary, the authors were able to study reaction times from 0.1 to 136 s with ∼100 ms time resolution, despite requiring ∼3 min to acquire laboratory SAXS curves. The authors followed-up this work by using the same flow system to study silver258 and palladium259 NP synthesis.
image file: d4lc00637b-f12.tif
Fig. 12 Laboratory-based X-ray scattering analysis. (a) A flow system for operando SAXS studies of the synthesis of gold nanoparticles. Different tubing lengths (Δd) between a commercial micro-mixer and the X-ray beam give access to different time points (used with permission from Polte et al., 2010; Copyright 2010 American Chemical Society).257 (b) A flow system for in situ SAXS studies of the polymerization-induced self-assembly comprising a heated reactor module and a cold-quench to stop the reaction before SAXS analysis. (c) Time-resolved SAXS data of nano-objects obtained with the system in (b) at reaction times controlled by the applied flow rate (adapted with permission from Guild et al. 2023; Copyright 2023 The Authors).268

A similar approach was later employed by Chen et al. and Herbst et al. to study the nucleation and growth of gold and ZnO NPs, respectively, over hours by simultaneous SAXS/WAXS and UV-vis spectroscopy.260,261 In contrast to Polte et al., once a steady flow was achieved within the capillary, flows were stopped and the reactions were followed within the static solutions. The authors analyzed the syntheses with a time resolution of between 0.5 and 10 minutes depending on the required acquisition time. Tillier et al. used a modified version of the electrochemical flow cell of Binninger et al. (reviewed in section 4.2.4) to study the degradation of carbon-supported Pt NP catalysts.262 Compared to their synchrotron experiments with exposure times on the second to minute time-scale, the acquisition of laboratory SAXS patterns with good signal-to-noise ratio required approximately 1 hour. Therefore, it is clear that depending on the kinetics of a reaction or process of interest, the experiment must be designed to consider the acquisition time required for lower flux laboratory sources.

6.1.2 Biological applications. Laboratory sources have also been utilized for biological applications. Bucciarelli et al. used a laboratory SAXS platform with an integrated millifluidic flow system to study biomolecules fractionated by inline size-exclusion chromatography (SEC).263 This so-called “SEC-SAXS” technique is utilized to produce monodisperse populations of often heterogeneous and unstable biological samples so that quantitative information on the structure, dynamics, and molecular weight of specific biomolecules can be extracted from SAXS data. This was the first time SEC-SAXS was performed without the use of synchrotron radiation, and the authors demonstrated that high quality data could be obtained with a similar amount of sample consumption and total experimental time with no radiation damage. Anaraki et al. used millifluidics and laboratory SAXS to study the stability of NP suspensions in the presence of biomolecules, an important parameter related to the behavior of nanoparticles in the environment and within the human body.264 The authors utilized a micromixer coupled to a quartz capillary to study the aggregation of silica NPs in the presence of human serum albumin (HSA) and in media with differing pH and ionic strength. They found that HSA significantly accelerates the aggregation of silica compared to the pH or ionic strength jumps studied.
6.1.3 Multi-source studies. Several research teams have also worked on projects using a combination of synchrotron radiation and laboratory sources. For example, Besenhard et al. used synchrotron PXRD (see section 3.3.3) and laboratory SAXS to study the synthesis of iron oxide nanoparticles.127 The authors used a continuous flow millifluidic device to assess the aggregation state of freshly precipitated NPs, finding that they were already highly aggregated after 5 s of reaction time. They subsequently used a semi-batch setup to study the de-agglomeration of the NPs using a neutralizing solution of citric acid, confirming de-agglomeration began within 20 s of citric acid addition. Lange et al. performed a synchrotron SAXS and WAXS study (see section 3.3.2) and used a laboratory SAXS platform to design and evaluate their microfluidic device before their beamtime.92 They evaluated the scattering profile of their chosen device material, OSTE+, as a potential X-ray window against more the commonly used Kapton and found that it had lower background and structure in the SAXS region. After selecting to use a completely OSTE+ device, they checked its transmission profile as a function of its thickness. They also refined the fabrication protocol to produce devices that would not age under the beam, and thus result in inconsistent background signal. Similarly, Monnier et al. used laboratory micro-XRD to complement their synchrotron XAS experiments on the corrosion of iron phases (see section 4.2.1).163,164 Micro-XRD was used to confirm the phase composition before and after redox processes and identify reaction intermediates that contributed to shifting absorption edges in the XANES spectra.

García-Lojo et al. performed a synchrotron and laboratory SAXS study of the assembly of NPs into colloidal supercrystals by pervaporation in microfluidic channels.265 In comparing 2D SAXS data collected at the synchrotron and the laboratory, they observed that single-crystal small angle diffraction spots were much better resolved in the synchrotron data due to the smaller beam size and higher angular resolution. Finally, Ehm et al. 3D printed a millifluidic dialysis chamber with COC windows for monitoring 100 μL samples by SAXS.266 They demonstrated the device by following reversible structural transitions in lipids and polymers induced by changing media pH and salt concentration over hours using acquisition times of 20 min. The in situ laboratory SAXS data were compared to ex situ data collected using synchrotron radiation with good agreement.

6.1.4 Exploring the limits of laboratory analysis. Two recent studies sought to explore the limits of in situ laboratory-based X-ray scattering analysis of micro- and milli-fluidic sample environments. Levenstein et al. studied a range of different samples, fluidic devices, and X-ray instruments and compared the type of data that could be obtained with an instrument optimized for performing SAXS/WAXS to an instrument optimized for performing single-crystal XRD.15 In particular, they investigated the potential of using a multi-frame measurement approach with short acquisition times to study processes with fast kinetics. They found that on the highly collimated SAXS/WAXS setup with lower flux at the sample (3.7 × 106 photons s−1), decent quality SAXS data from flowing silica NP solutions could be obtained with even sub-second acquisition times. However, the WAXS signal-to-noise ratio from such short acquisitions was too low to obtain valuable data. Conversely, on the microfocused XRD setup with higher flux at the sample (5.7 × 109 photons s−1), the authors could obtain powder diffraction patterns from flowing organic and inorganic crystallites with acquisition times as short as 25 ms. These results demonstrate that WAXS/XRD analysis is clearly flux limited on laboratory sources, yet SAXS analysis, which requires more collimated beams to avoid smearing at small angles, appears less restricted by the source flux.

Radajewski et al. performed a microfluidic SAXS study on a state-of-the-art laboratory SAXS platform that was modified to obtain a small 0.25 × 0.25 mm2 beam while providing a flux density of almost 107 photons s−1.267 In particular, they investigated the quality of data that could be obtained from flows of strongly scattering inorganic nanoparticles to moderately to weakly scattering proteins and polymer materials. For the high contrast gold and silica NPs studied, 10 min exposures were enough to obtain good quality SAXS patterns that could be fit with form and structure factors and for which the invariant could be calculated. For more weakly scattering bovine serum albumin (BSA) and latex NPs, longer 1 hour acquisitions were required to fit form factors, and even in these cases, increased noise especially at high q, made it difficult to resolve all the structure peaks of dilute samples. However, these results are very promising, especially for stronger contrast inorganic materials, and the use of a rotating anode or liquid metal jet source could provide between one and two orders of magnitude greater flux with the same optical configuration.

6.1.5 Laboratory facilities dedicated to in situ measurements. In the previous subsections, we highlighted research from several groups who have developed methodology for in situ laboratory-based X-ray scattering analysis including at the Federal Institute for Materials Research and Testing (BAM, DE),257,258 the Förster group (DE),260,261 and our own laboratory at CEA Paris-Saclay (FR).92,254,290 These advances have led to the recent creation of two new laboratory X-ray facilities dedicated to operando X-ray experiments. The first is the DL-SAXS facility managed by the I22 beamline of Diamond Light Source (UK).268 This facility comprises a commercial SAXS/WAXS platform with a high flux liquid metal jet source, and it is designed to enable tests of sample environments before use on the beamline or as an alternative to synchrotron radiation. Guild et al., recently demonstrated the use of this facility to study the self-assembly of block co-polymer nano-objects with a millifluidic continuous flow system (Fig. 12b).268 Using acquisition times of 5 min, the authors were able to study the polymer self-assembly process over reaction times of 1 to 25 minutes controlled by adjusting the flow rate of reactants through the system and into the analysis capillary. These measurements revealed a two-step assembly process, beginning with the formation of loose hydrated aggregates between 60 and 100 s and the subsequent densification of the aggregates into well-defined spherical nano-objects over ∼300–750 s (Fig. 12c).

The second platform is the Flow-Xl National Facility for Analysis of Crystallization in Flow Systems located at the University of Leeds (UK).269 Flow-Xl is built around an X-ray diffractometer with a microfocused rotating anode source that is also coupled to a Raman spectrometer for performing simultaneous XRD/Raman. This facility was designed specifically for performing operando flow-based experiments, and as such, is equipped with a range of sample environments from millifluidic flow cells and humidity chambers to microfluidic devices. Commissioning experiments on this platform performed by Turner et al. demonstrated its potential for studying the nucleation and growth of inorganic and organic materials from aqueous solution.269 XRD and Raman yielded information on the dynamics of both the solid phases and solution chemistry, with limits of detection of 0.02–0.1 wt% and 0.625–2.5 g L−1, respectively, depending on the materials and solution species studied. For operando analysis of the cooling crystallization of NaSO4, XRD and Raman acquisition times of 18 and 17 s, respectively, were found to be sufficient to obtain data with good signal-to-noise ratio.

6.2 X-ray spectroscopy

6.2.1 X-ray absorption spectroscopy. Less work has been performed on micro- and milli-fluidic laboratory-based X-ray spectroscopy compared to X-ray scattering and diffraction. Taking XAS as an example, even for ex situ experiments, it is still most often performed at synchrotron radiation facilities.291 While laboratory-based XAS is indeed becoming more common, most in situ studies to date have utilized Norby-type capillary devices (see section 2.1.2).292,293 Despite this, we are aware of at least one millifluidic laboratory XAS study. Kozyr et al. developed a 3D-printed photocatalytic cell for performing XANES with inline mass spectrometry of gas production.270 This cell had two windows, one for holding a TiO2 support film and performing XANES in fluorescence mode and the other for UV irradiation. The cell also had a fluid port for the injection of a Pt salt solution and a gas inlet and outlet for applying a continuous flow of Ar gas. The authors used this cell to follow the photodeposition of a Pt catalyst onto the support and its subsequent catalytic activity in a hydrogen evolution reaction. Although due to the low flux of the laboratory source, the acquisition of full XANES spectra was too slow for operando kinetics studies. Therefore, only three representative wavelengths (pre-edge, edge, and post-edge) were selected to follow with time.
6.2.2 X-ray fluorescence. XRF instruments are much more commonly found in laboratories than XAS instruments, and perhaps for this reason, there exist a few more micro- and milli-fluidic laboratory XRF studies. To the best of our knowledge, the first example comes from Miller et al., who coupled a classic microfluidic capillary electrophoresis device to a benchtop micro-XRF spectrometer.271 The authors demonstrated the separation and detection of a variety of metal and organic species with a comparable limit of detection (∼10−4 M) to previous synchrotron studies.151,152 Greaves and Manz built a lead-lined miniaturized XRF device containing a small radioactive Am source for on-chip production of X-rays.18 They analyzed the sensitivity of the device to detect various metal films inserted into the chip and obtained decent signal-to-noise ratios despite the relatively low flux of the Am source compared to standard XRF instruments. More recently, Maurice et al. reported an automated setup for microfluidic liquid–liquid extraction of metal waste streams using inline XRF (Fig. 13).272 They demonstrated the utility of the system by performing extractions of three rare earth elements and quantifying the metal concentration within both liquid phases as a function of the fluid contact time. Olivier et al. followed-up this study by developing a similar millifluidic platform with inline XRF for quantifying solid-phase metal extraction.273 There are additional examples of micro- and milli-fluidic laboratory-based XRF studies where 2D or 3D spatial mapping was performed, and these will be covered in the next subsection.
image file: d4lc00637b-f13.tif
Fig. 13 Experimental setup for laboratory-based X-ray fluorescence analysis (left). (i) View of the measurement area. (ii) Illustration of the microchannel path and the elliptical footprint of the X-ray beam. (iii) Cross-section of the microchannel. (iv) Photograph of the XRF device (used with permission from Maurice et al.; Copyright 2021 The Authors).272

6.3 X-ray imaging

6.3.1 Development of laboratory-based in situ μCT. Similar to X-ray scattering and diffraction, there has also been a great deal of work in micro- and milli-fluidic devices for in situ and operando X-ray imaging with laboratory sources. Most of this has been focused on X-ray tomography as the small beam sizes and soft X-rays normally utilized for STXM are not readily available in the laboratory and decreased flux makes scanning techniques impractical for time-resolved experiments. Before the development of miniaturized flow cells for laboratory μCT, the wetting and diffusion dynamics within cm-scale rock samples were first analyzed in 2D with X-ray radiography294 and then in 3D using X-ray tomography.295 Various groups also developed new laboratory X-ray imaging instruments for visualizing flows, such as X-ray stereography setups with two perpendicular source-detector pairs for obtaining 3D information without requiring the slow tomographic rotation of a sample.296 Work was also done in enabling helical-CT, where a sample is both rotated and translated vertically during tomogram acquisition to reduce beam artefacts and measure longer samples,297,298 and in using gantry-based systems where the source and detector are rotated instead of the sample.298 Although this list is far from exhaustive, we note in particular developments at the Center for Multiphase Flow Research and Education at Iowa State University (CoMFRE, US),296 the Micro CT Facility of the Australian National University (CTLab),297,299 the Ghent University Centre for X-ray Tomography (UGCT, BE),298 and the Henry Moseley X-Ray Imaging Facility at the University of Manchester (HMXIF, UK)280 for improving in situ μCT of fluid flow.
6.3.2 μCT for geophysical studies. To the best of our knowledge, Youssef et al. performed the first operando flow-based laboratory μCT study.274 They used a millifluidic Hassler-type cell made from PEEK to analyze an oil–water drainage/imbibition cycle within a 6 mm-diameter core of sandstone. The authors used different oil flow rates that produced equilibrium conditions within the core and visualized the various 3D fluid distributions that developed. Continued flow ensured that the fluid distribution remained relatively stationary, so that even long tomogram acquisition times of 80 minutes could reliably capture different steps in the drainage/imbibition cycle. Andrew et al. performed a similar experiment using a high-pressure Hassler flow cell to study geological CO2 sequestration in limestone.275 In this experiment, 2D projections were collected continuously during the injection of super-critical CO2 until reaching steady-state, at which point a full tomogram was acquired. The authors later expanded their study to analyze the trapping of super-critical CO2 in a range of different rocks276 and worked to quantify wetting in the three-phase CO2-brine-limestone system by extracting contact angles from 2D slices of tomograms.277 Bultreys et al. performed the first laboratory-based fast-μCT study of flow within porous media.278 Using a gantry-based system with a microfocus X-ray source and CMOS detector, the authors were able to acquire full tomograms within 12 s with a voxel size of ∼15 μm. Data quality at this fast speed was improved using information on the sample obtained a priori from a higher resolution tomogram collected under equilibrium conditions.
6.3.3 μCT for geochemical studies. As for synchrotron-based studies, laboratory-based μCT studies have also investigated geochemical processes occurring within porous media. For example, Menke et al. used a millifluidic Hassler cell to study the dissolution of a limestone core during the injection of a CO2-saturated brine at 50 °C and 10 MPa.279 They followed the dissolution of the pores over ∼2.5 hours with a spatial and temporal resolution of 3.8 μm and 15 minutes, respectively. Gajjar et al. studied the precipitation of barite within a microporous glass column.280 The authors used an innovative approach in which projections were acquired at angles calculated by the golden-ratio rather than sampling at equally spaced angles around the sample. While taking longer to acquire each projection, this approach required fewer projections to reconstruct a tomogram. In this way, the authors tried to find a protocol that optimized the spatial and temporal resolution of tomograms. Godinho et al. used a similar microporous column to study calcite precipitation and performed the first operando helical-CT study of flow in porous media.281 They followed the slow growth of calcite over 17 days with nine tomograms collected at different times, each taking 2.5 hours. More recently, Singh et al. performed a μCT study of calcite dissolution with a tomogram acquisition time of 83 s and a voxel size of 19.3 μm.282
6.3.4 Tomographic X-ray PIV/PTV. Before 3D tomographic X-ray particle image velocimetry (PIV) or particle tracking velocimetry (PTV) was performed at the synchrotron, at least two groups implemented it with a laboratory source. Mäkiharju et al. performed a proof-of-concept study by analyzing a creeping flow of 60 μm diameter tracer particles within a single-phase pipe flow, a microporous glass column, and a thin liquid film around a flow of gas slugs (Fig. 14).283 In each case, the cylindrical flow channel was continuously rotated at ∼20 rpm while projections were acquired at 68 Hz (14.5 ms per projection) and full tomograms completed in ∼3 seconds. As the tracer particles were moving at only ∼20–50 μm s−1 in the pipe flow, the authors could collect high contrast projections with negligible blur and use standard PIV/PTV algorithms to obtain the 3D velocity fields. However, the flow constrictions in the other two cases increased the flow velocity, so that only particles that had become stuck could be resolved. Bultreys et al. performed a similar analysis of a creeping flow within porous media and successfully resolved the flowing particles to extract the 3D velocity field.284 While they used a slower scan speed of 100 ms per projection and 70 s for acquiring a full tomogram, they utilized flow rates of only tens of nanoliters per minute that produced particle velocities of ∼100–200 nm s−1. They also utilized a filtered backprojection algorithm during reconstruction that interleaved data between two consecutive tomograms, effectively decreasing the time step between tomograms to 35 s. Together, the flow rates and reconstruction methods meant that, on average, particles moved only 0.5 voxels between frames providing a sufficient resolution for particle tracking. Despite this progress, it is clear that laboratory-based 3D X-ray PIV can only follow slow flow dynamics, and its time resolution cannot compare to subsequent work in synchrotron-based X-ray PIV (see section 5.3.2).
image file: d4lc00637b-f14.tif
Fig. 14 Experimental setup for laboratory-based 3D tomographic X-ray particle tracking velocimetry of a creeping millifluidic flow (used with permission from Mäkiharju et al.; Copyright 2021 The Authors.283
6.3.5 XRF mapping. There has been some limited work in laboratory-based μXRF for in situ microfluidic chemical mapping. Tsuji et al. developed microfluidic chips for concentrating medical and environmental samples for elemental analysis with a benchtop XRF spectrometer.285 They developed two chips, one for preparing samples for grazing-incidence XRF and one for scanning μXRF. In the case of μXRF, the authors demonstrated that a nL sample of solution containing just nanograms of Cu, Cd, and Fe could be concentrated, dried, and then mapped on-chip. Subsequently, Nakano and Tsuji performed confocal 3D-XRF on a multi-layer polymer microfluidic device containing separate microchannels within the different layers.286 To demonstrate the method, they filled one channel with a Cu solution and one with a Co solution (both 3 mg mL−1) and obtained a 3D map that distinguished both channels with elemental selectivity. Finally, McIntosh et al. developed a polycarbonate-Kapton microfluidic device for monitoring Pu concentrations in nuclear waste.287 Using Sr as a surrogate for Pu due to its similar fluorescent signature, the authors demonstrated that chips could be filled directly from a micropipette with 5 μL of solution and subsequently mapped for their Sr content with a limit of detection of ∼5 ppm.

7 Discussion and outlook

The numerous examples cited in this review demonstrate the huge interest in using micro/millifluidics to carry out in situ and operando X-ray analyzes in both liquids and gases. We have seen that there is interest across all types of X-ray analysis, whether scattering, spectroscopy, or imaging and using both synchrotron and laboratory sources. In total, we have reviewed 148 articles (139 explicitly found in Tables 1–4), which we believe represents a near-comprehensive account of micro- and milli-fluidic X-ray literature in the physical sciences from the year 1999 until 2023. The growth in the number of articles produced is clear (Fig. 15), where the average number of papers produced per year from 1999 to 2015 was ∼4 and the number per year since the life sciences review of Ghazal et al.12 in 2016 is ∼10.
image file: d4lc00637b-f15.tif
Fig. 15 Chronology of the articles covered in this review (excluding articles from 2024).

There is good reason for this high and growing level of interest. Specifically for continuous flow devices, the renewal of fluid under the beam has several advantages for X-ray measurements, including increasing their temporal resolution. In continuous flow, time resolution does not depend on the frame rate of a detector, but rather on the time it takes for fluid to pass through the beam and, therefore, on the linear velocity of the fluid as well as the beam size (and, in certain cases, on the initial mixing processes). Inline fluidic devices also make it possible to reduce the aging time of the sample before analysis by eliminating the step of introducing the sample under beam. This both reduces the observation dead time while guaranteeing the repeatability of measurements, e.g., by eliminating sample preparation artefacts. Using an inline fluidic configuration can also increase the rate at which samples can be measured thanks to the possibility of changing the nature of the sample (e.g., flow rate ratio, temperature) without a disassembly/reassembly operation, a particularly time-consuming process for techniques requiring operation under vacuum. This time saving makes it possible to accelerate data collection and even consider performing large screening studies. Finally, the continuous renewal of the sample under an ionizing beam limits the effects of radiation damage.

The size reduction associated with microfluidic devices adds further advantages to the previous list. It makes it possible to reduce the necessary sample volumes by one or more orders of magnitude, to carry out transmission analysis of highly absorbing samples (crucial in the soft X-ray domain), to access reaction times of less than a millisecond due to very efficient mixing functions at the micro-scale, and more generally, to combine many types of sample manipulation (e.g., droplet generation, separation, heating) into a single platform. It also results in an increase in the surface area-to-volume ratio, which can be a significant advantage when the objective is to probe the interactions between the substrate of the device and the objects conveyed by the flow, for instance, in catalysis studies. This increase of surface effects can, however, also constitute a point of weakness when uncontrolled events of adsorption and/or heterogeneous nucleation on the walls of a microchannel obscure or alter the bulk phenomena of interest.102

Initially condensing into a field of research in its own right,96 microfluidics quickly became a formidable tool for manipulating samples in domains as varied as biology, clinical diagnostics, synthetic chemistry, and materials science.83,84,300,301 Many research groups and companies have found success in applying microfluidic tools to these various areas. However, “going microfluidic” normally requires a large investment in equipment and technical skill, and this has been no different for groups looking to use fluidic devices for in situ X-ray analysis. In our opinion, perhaps the greatest bottleneck here has been in the complex microfabrication techniques and specific environments required for making conventional microfluidic devices. Such environments, of the “cleanroom” type, are characterized by a very low concentration of particles in the air as well as controlled temperature and relative humidity. Various specific equipment, such as spin-coaters, UV mask aligners, and plasma cleaners are also necessary. The investment, both in terms of cost and manpower, needed to establish such an environment cannot, therefore, be justified when its use is only occasional. This strongly hinders the uptake of microfluidics by non-specialist laboratories.

Fortunately, this barrier is continually being lowered due to the increasing availability of commercial solutions and the development of more accessible methods in microfabrication. Some companies (e.g., Microfluidic ChipShop, Micronit, Protochips) offer turnkey microfluidic devices associated with different fluidic functions such as droplet production, mixing, and electrochemistry. This solution thus relieves the user of any microfabrication duties and might only require access to a laminar flow hood to guarantee sufficient cleanliness when setting up the fluidic system. Indeed, many of the studies reviewed herein were performed with commercially available components or completely commercial chips, for example, ref. 160, 165 and 287 and several others.

At the same time, microfabrication is beginning to require less expertise due to new technologies that reduce the number of fabrication steps and guarantee very good device reproducibility and reliability. For photolithographic cleanroom-based fabrication, the use of commercial polymer films (i.e., dry film photoresists) of calibrated thickness for the creation of molds avoids the production of photoresist films by spin-coating, the result of which strongly depends on the temperature and humidity of the laboratory.302 Many non-cleanroom routes are also available. The use of hot embossable materials such as cyclic olefin copolymers (e.g., TOPAS®) or cyclic olefin polymers (e.g., ZEONOR®, ZEONEX®) makes it possible to replicate the same device many times in an automated and repeatable manner.303,304 The use of laser micromachining makes it possible to produce monolithic devices from many types of materials and/or to produce “sandwich”-type devices combining different materials, again in an automated manner.102,305,306 The main challenge with the high-performing femtosecond laser machining platforms required for patterning small features in a large range of materials is that their cost is on the scale of large cleanroom equipment. Much less expensive, 3D printing now makes it possible to directly print devices with minimum channel sizes below 100 μm using commercial printers with minimal to no modifications.125,307 Indeed, a large number of the studies used here were performed with 3D-printed components or chips (e.g., ref. 131, 216 and 266). Further, the adoption of simple plug-and-play millifluidic devices or millifluidic devices with lower requirements in manufacturing precision can still produce excellent data (see ref. 127, 262 and 273 for example). The multiplicity of available fabrication techniques now enables one to be more selective in the choice of device material to optimize key parameters such as its X-ray absorption or scattering profile or even its degradation under irradiation. For example, in our laboratory, where we perform a variety of different X-ray techniques for different types of samples, we have investigated every process mentioned above and found them all to be useful depending on the technique requirements and demands of the particular experiment.

These developments should ultimately encourage the use of micro- and milli-fluidic sample environments for in situ characterization at the synchrotron and in the laboratory. Until very recently, synchrotron studies conducted with microfluidic devices produced relatively little data due to the numerous technical challenges encountered during their implementation. Their handling, already delicate, is made even more difficult at the synchrotron and away from one's own laboratory environment. Consequently, journal publications only reflect a small part of the real effort made by the scientific community to develop micro- and milli-fluidic tools that better exploit the potential of X-ray analysis. For their part, synchrotron facilities have largely participated in this effort by developing microfocused beams compatible with the typical size of microfluidic channels (typically hundreds of microns). Many synchrotron facilities have also formed dedicated teams to help users with sample preparation in general, or microfluidics in particular, such as the Partnership for Soft Condensed Matter at the ESRF,125 the Microfluidic Laboratory at SOLEIL,227 and the Sample Environment Development Laboratory at Diamond Light Source.

Additional actions on their part could further facilitate micro- and milli-fluidic experiments. These include the installation of “clean” zones in the form of laminar flow hoods or biosafety cabinets (ISO 7 or 5)—essential for the preparation of the devices before experiments—near the beamline experimental hutch, the provision of commercial microfluidic environments already tested and integrated by the beamline staff, and the inclusion of inline optical microscopes for the visualization of microfluidic channels and the identification of measurement points. To our knowledge, automated or remote access to microfluidic equipment has not yet been implemented at a synchrotron beamline. However, this has been demonstrated for stopped-flow devices at bioSAXS beamlines (e.g., SSRL 4.2, PETRA III P12) and may be possible in the future for continuous flow micro- and milli-fluidic platforms for greater accessibility.

There are also several things beamline users can do to increase the success rate of their microfluidics beamtimes. The first is ensuring that the devices they bring are safe and reliable, e.g., not prone to leaking on expensive beamline hardware. Beamline staff have invested significant time in building their unique instruments and must continue to devote time to hosting user experiments. When users come with unreliable devices that do not produce data and thus, do not produce papers, they are less enthusiastic about recommending future microfluidics experiments. We have seen above that there are many routes to making robust micro- and milli-fluidic devices. We have also seen in section 6 that laboratory-based analysis is becoming more and more feasible. For this reason, we recommend that potential users try to perform an experiment with a laboratory instrument before applying for synchrotron beamtime. For example, the Sample Environment Development Laboratory at Diamond even offers this opportunity in partnership with beamline I22 and their offline SAXS platform.268 Even if an experiment is not feasible with the lower flux of a laboratory source, using a laboratory instrument to first characterize the scattering/absorption of a device such as done by Lange et al.92 or practicing the mounting of a device to avoid any potential surprises at a beamtime can go a long way.

X-ray analysis techniques, whether performed with a synchrotron or a laboratory source, are powerful tools for structural and chemical characterization. When combined with a micro- or milli-fluidic sample environment, they offer unparalleled possibilities for in situ and operando studies. The combination makes it possible, on one hand, to eliminate the preparation artifacts inherent in post-mortem analyses, and substantially improves, on the other hand, the temporal resolution compared to studies in a static configuration (i.e., without fluid renewal). However, while synchrotron light sources have long been and will continue to be a powerful tool in the arsenal of crystallographers and chemists, the relative inaccessibility of these facilities does limit the progress of individual research teams, which could otherwise measure more samples, develop more robust workflows, and iterate more rapidly. It is here we see tremendous potential for the current generation of commercial laboratory X-ray instruments, not just to permit more efficient use of synchrotron beamtime, but also to support completely independent in situ flow-based analysis platforms (see section 6.1.5).

Perhaps the most innovative prospect in the use of micro- and milli-fluidic X-ray sample environments is the possibility of exercising feedback control over the injected reagents, temperature, or other reaction conditions. Real-time analysis of data coupled with machine learning, as demonstrated by Fong et al.111 and Younes et al.112 for SAXS (section 3.2.3), makes it possible to direct the system towards a state corresponding to a pre-defined structural or chemical parameter. For example, in the field of (nano)materials synthesis, this ability should facilitate a departure from the classical trial-and-error approach and guarantee both the outcome of the syntheses and their reproducibility. Another recent study reported an autonomous laboratory for the high temperature synthesis of solid powders and their characterization by XRD.308 We could envision a similar type of system,309 but one using XRD/SAXS analysis for automated liquid phase synthesis with a micro- or milli-fluidic device benefiting from all of the advantages discussed above. The industrial value of such a paradigm shift would be considerable. Thus, making these techniques more accessible to the point of being routine, both at the synchrotron and in the laboratory, is a worthwhile effort not only for gaining fundamental scientific understanding, but also for developing new materials and processes for industrial and societal use.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Author contributions

MAL conceived the idea for the review and refined the concept in discussions with all authors. MAL performed the literature survey for sections 2, 3, 4, and 6. MAL and CC performed the literature survey for section 5. MAL wrote the manuscript with contributions from CC. MAL and CC edited the manuscript. FM, FT, and OT provided feedback in their areas of expertise.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work received state support managed by the French National Research Agency (ANR) under the France 2030 framework (ANR-22-PEXD-0006 and ANR-22-PEXD-0007). The authors would like to thank the many beamline scientists and staff who have supported their micro- and milli-fluidics beamtimes over the years, with a special mention of Sarah Day, Chiu Tang, Manfred Burghammer, Javier Perez, Thomas Bizien, Stefan Stanescu, Rachid Belkhou, Diego Pontoni, and Peter van der Linden. We also thank our editor, David Lake, for his assistance and patience during the preparation of the review.

References

  1. J. J. De Yoreo, P. U. P. A. Gilbert, N. A. J. M. Sommerdijk, R. L. Penn, S. Whitelam, D. Joester, H. Zhang, J. D. Rimer, A. Navrotsky, J. F. Banfield, A. F. Wallace, F. M. Michel, F. C. Meldrum, H. Cölfen and P. M. Dove, Science, 2015, 349, aaa6760 CrossRef.
  2. A. E. S. Van Driessche, N. Van Gerven, P. H. H. Bomans, R. R. M. Joosten, H. Friedrich, D. Gil-Carton, N. Sommerdijk and M. Sleutel, Nature, 2018, 556, 89–94 CrossRef CAS.
  3. Z. Gao, M. Odstrcil, S. Böcklein, D. Palagin, M. Holler, D. F. Sanchez, F. Krumeich, A. Menzel, M. Stampanoni, G. Mestl, J. A. van Bokhoven, M. Guizar-Sicairos and J. Ihli, Sci. Adv., 2021, 7, eabf6971 CrossRef CAS.
  4. M. Macino, A. J. Barnes, S. M. Althahban, R. Qu, E. K. Gibson, D. J. Morgan, S. J. Freakley, N. Dimitratos, C. J. Kiely, X. Gao, A. M. Beale, D. Bethell, Q. He, M. Sankar and G. J. Hutchings, Nat. Catal., 2019, 2, 873–881 CrossRef CAS.
  5. D. P. Finegan, M. Scheel, J. B. Robinson, B. Tjaden, I. Hunt, T. J. Mason, J. Millichamp, M. Di Michiel, G. J. Offer, G. Hinds, D. J. L. Brett and P. R. Shearing, Nat. Commun., 2015, 6, 6924 CrossRef CAS.
  6. S. N. S. Hapuarachchi, Z. Sun and C. Yan, Adv. Sustainable Syst., 2018, 2, 1700182 CrossRef.
  7. Z. Ning, D. S. Jolly, G. Li, R. De Meyere, S. D. Pu, Y. Chen, J. Kasemchainan, J. Ihli, C. Gong, B. Liu, D. L. R. Melvin, A. Bonnin, O. Magdysyuk, P. Adamson, G. O. Hartley, C. W. Monroe, T. J. Marrow and P. G. Bruce, Nat. Mater., 2021, 20, 1121–1129 CrossRef CAS.
  8. D. H. Bilderback, P. Elleaume and E. Weckert, J. Phys. B, 2005, 38, S773 CrossRef CAS.
  9. H. Weise and W. Decking, presented in part at the 38th International Free Electron Laser Conference, FEL2017, Santa Fe, NM, USA, 2017 Search PubMed.
  10. S. Shin, AAPPS Bull., 2021, 31, 21 CrossRef.
  11. T. Skarzynski, Acta Crystallogr., Sect. D: Biol. Crystallogr., 2013, 69, 1283–1288 CrossRef CAS PubMed.
  12. A. Ghazal, J. P. Lafleur, K. Mortensen, J. P. Kutter, L. Arleth and G. V. Jensen, Lab Chip, 2016, 16, 4263–4295 RSC.
  13. M. A. Bañares, Catal. Today, 2005, 100, 71–77 CrossRef.
  14. M. Ilett, M. Afzali, B. Abdulkarim, Z. Aslam, S. Foster, M. Burgos-Ruiz, Y.-Y. Kim, F. C. Meldrum and R. M. Drummond-Brydson, J. Microsc., 2024, 1–14,  DOI:10.1111/jmi.13300.
  15. M. A. Levenstein, K. Robertson, T. D. Turner, L. Hunter, C. O'Brien, C. O'Shaughnessy, A. N. Kulak, P. Le Magueres, J. Wojciechowski, O. O. Mykhaylyk, N. Kapur and F. C. Meldrum, IUCrJ, 2022, 9, 538–543 CrossRef CAS.
  16. C. L. Hansen, E. Skordalakes, J. M. Berger and S. R. Quake, Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 16531–16536 CrossRef CAS.
  17. B. Zheng, J. D. Tice, L. S. Roach and R. F. Ismagilov, Angew. Chem., Int. Ed., 2004, 43, 2508–2511 CrossRef CAS.
  18. E. D. Greaves and A. Manz, Lab Chip, 2005, 5, 382–391 RSC.
  19. E. Lawrence Bright, C. Giacobbe and J. P. Wright, J. Synchrotron Radiat., 2021, 28, 1377–1385 CrossRef PubMed.
  20. U. Weierstall, Philos. Trans. R. Soc., B, 2014, 369, 20130337 CrossRef PubMed.
  21. S. Sui and S. L. Perry, Struct. Dyn., 2017, 4, 032202 CrossRef.
  22. L. Paulson, S. R. Narayanasamy, M. L. Shelby, M. Frank and M. Trebbin, Struct. Dyn., 2024, 11, 011101 CrossRef CAS.
  23. I. Grillo, Curr. Opin. Colloid Interface Sci., 2009, 14, 402–408 CrossRef CAS.
  24. J. T. Avaro, S. L. P. Wolf, K. Hauser and D. Gebauer, Angew. Chem., Int. Ed., 2020, 59, 6155–6159 CrossRef CAS PubMed.
  25. R. J. Clarke and M. A. A. Khalid, in Pumps, Channels, and Transporters, 2015, pp. 179–209,  DOI:10.1002/9781119085126.ch7.
  26. O. L. J. Virtanen, M. Kather, J. Meyer-Kirschner, A. Melle, A. Radulescu, J. Viell, A. Mitsos, A. Pich and W. Richtering, ACS Omega, 2019, 4, 3690–3699 CrossRef CAS.
  27. S. Schmölzer, D. Gräbner, M. Gradzielski and T. Narayanan, Phys. Rev. Lett., 2002, 88, 258301 CrossRef.
  28. G. Guilera, M. A. Newton, C. Polli, S. Pascarelli, M. Guinó and K. K. Hii, Chem. Commun., 2006, 41, 4306–4308 RSC.
  29. B. Abécassis, F. Testard, O. Spalla and P. Barboux, Nano Lett., 2007, 7, 1723–1727 CrossRef PubMed.
  30. J. Polte, T. T. Ahner, F. Delissen, S. Sokolov, F. Emmerling, A. F. Thünemann and R. Kraehnert, J. Am. Chem. Soc., 2010, 132, 1296–1301 CrossRef CAS.
  31. T. M. Stawski, A. E. S. van Driessche, M. Ossorio, J. Diego Rodriguez-Blanco, R. Besselink and L. G. Benning, Nat. Commun., 2016, 7, 11177 CrossRef CAS PubMed.
  32. M. L. Whittaker, P. J. M. Smeets, H. Asayesh-Ardakani, R. Shahbazian-Yassar and D. Joester, Angew. Chem., Int. Ed., 2017, 56, 16028–16031 CrossRef CAS.
  33. J. Bolze, B. Peng, N. Dingenouts, P. Panine, T. Narayanan and M. Ballauff, Langmuir, 2002, 18, 8364–8369 CrossRef CAS.
  34. J. Cravillon, C. A. Schroder, R. Nayuk, J. Gummel, K. Huber and M. Wiebcke, Angew. Chem., Int. Ed., 2011, 50, 8067–8071 CrossRef CAS PubMed.
  35. F. Carraro, J. D. Williams, M. Linares-Moreau, C. Parise, W. Liang, H. Amenitsch, C. Doonan, C. O. Kappe and P. Falcaro, Angew. Chem., Int. Ed., 2020, 59, 8123 CrossRef CAS PubMed.
  36. M. J. Quayle, R. J. Davey, A. J. McDermott, G. J. T. Tiddy, D. T. Clarke and G. R. Jones, Phys. Chem. Chem. Phys., 2002, 4, 416–418 RSC.
  37. P. J. Chupas, K. W. Chapman, C. A. Kurtz, J. C. Hanson, P. L. Lee and C. P. Grey, J. Appl. Crystallogr., 2008, 41, 822–824 CrossRef CAS.
  38. C. G. Roessler, R. Agarwal, M. Allaire, R. Alonso-Mori, B. Andi, J. F. R. Bachega, M. Bommer, A. S. Brewster, M. C. Browne, R. Chatterjee, E. Cho, A. E. Cohen, M. Cowan, S. Datwani, V. L. Davidson, J. Defever, B. Eaton, R. Ellson, Y. Feng, L. P. Ghislain, J. M. Glownia, G. Han, J. Hattne, J. Hellmich, A. Héroux, M. Ibrahim, J. Kern, A. Kuczewski, H. T. Lemke, P. Liu, L. Majlof, W. M. McClintock, S. Myers, S. Nelsen, J. Olechno, A. M. Orville, N. K. Sauter, A. S. Soares, S. M. Soltis, H. Song, R. G. Stearns, R. Tran, Y. Tsai, M. Uervirojnangkoorn, C. M. Wilmot, V. Yachandra, J. Yano, E. T. Yukl, D. Zhu and A. Zouni, Structure, 2016, 24, 631–640 CrossRef CAS PubMed.
  39. R. H. Morris, E. R. Dye, D. Axford, M. I. Newton, J. H. Beale and P. T. Docker, Sci. Rep., 2019, 9, 12431 CrossRef CAS PubMed.
  40. B. Chance, Rev. Sci. Instrum., 1951, 22, 619–627 CrossRef CAS.
  41. R. L. Berger, B. Balko and H. F. Chapman, Rev. Sci. Instrum., 1968, 39, 493–498 CrossRef CAS PubMed.
  42. P. Panine, S. Finet, T. M. Weiss and T. Narayanan, Adv. Colloid Interface Sci., 2006, 127, 9–18 CrossRef CAS PubMed.
  43. L. Matthews and T. Narayanan, Colloid Polym. Sci., 2023, 301, 721–728 CrossRef CAS.
  44. A. Zabilska, A. H. Clark, D. Ferri, M. Nachtegaal, O. Kröcher and O. V. Safonova, Phys. Chem. Chem. Phys., 2022, 24, 21916–21926 RSC.
  45. C. K. Christensen, M. A. Karlsen, A. O. Drejer, B. P. Andersen, C. L. Jakobsen, M. Johansen, D. R. Sorensen, I. Kantor, M. R. V. Jorgensen and D. B. Ravnsbaek, J. Synchrotron Radiat., 2023, 30, 561–570 CrossRef CAS.
  46. P. Norby, Mater. Sci. Forum, 1996, 228–231, 147–152 CAS.
  47. R. Hodyss, T. H. Vu, M. Choukroun and M. L. Cable, J. Appl. Crystallogr., 2021, 54, 1268–1270 CrossRef CAS.
  48. B. S. Clausen, G. Steffensen, B. Fabius, J. Villadsen, R. Feidenhans'l and H. Topsøe, J. Catal., 1991, 132, 524–535 CrossRef CAS.
  49. P. Norby, J. Am. Chem. Soc., 1997, 119, 5215–5221 CrossRef CAS.
  50. K. M. O. Jensen, H. L. Andersen, C. Tyrsted, E. D. Bojesen, A. C. Dippel, N. Lock, S. J. L. Billinge, B. B. Iversen and M. Christensen, ACS Nano, 2014, 8, 10704–10714 CrossRef CAS PubMed.
  51. J. A. Rodriguez, X. Wang, P. Liu, W. Wen, J. C. Hanson, J. Hrbek, M. Pérez and J. Evans, Top. Catal., 2007, 44, 73–81 CrossRef CAS.
  52. T. R. Jensen, T. K. Nielsen, Y. Filinchuk, J.-E. Jorgensen, Y. Cerenius, E. M. Gray and C. J. Webb, J. Appl. Crystallogr., 2010, 43, 1456–1463 CrossRef CAS.
  53. N. V. Y. Scarlett, D. Hewish, R. Pattel and N. A. S. Webster, Rev. Sci. Instrum., 2017, 88, 105104 CrossRef PubMed.
  54. T.-D. Nguyen-Phan, Z. Liu, S. Luo, A. D. Gamalski, D. Vovchok, W. Xu, E. A. Stach, D. E. Polyansky, E. Fujita, J. A. Rodriguez and S. D. Senanayake, J. Phys. Chem. C, 2016, 120, 3472–3482 CrossRef CAS.
  55. J. Becker, M. Bremholm, C. Tyrsted, B. Pauw, K. M. O. Jensen, J. Eltzholt, M. Christensen and B. B. Iversen, J. Appl. Crystallogr., 2010, 43, 729–736 CrossRef CAS.
  56. P. Ferrer, I. da Silva, J. Rubio-Zuazo, B. F. Alfonso, C. Trobajo, S. Khainakov, J. R. Garcia, S. Garcia-Granda and G. R. Castro, J. Synchrotron Radiat., 2012, 19, 93–100 CrossRef CAS.
  57. H. N. Chapman, P. Fromme, A. Barty, T. A. White, R. A. Kirian, A. Aquila, M. S. Hunter, J. Schulz, D. P. DePonte, U. Weierstall, R. B. Doak, F. Maia, A. V. Martin, I. Schlichting, L. Lomb, N. Coppola, R. L. Shoeman, S. W. Epp, R. Hartmann, D. Rolles, A. Rudenko, L. Foucar, N. Kimmel, G. Weidenspointner, P. Holl, M. N. Liang, M. Barthelmess, C. Caleman, S. Boutet, M. J. Bogan, J. Krzywinski, C. Bostedt, S. Bajt, L. Gumprecht, B. Rudek, B. Erk, C. Schmidt, A. Homke, C. Reich, D. Pietschner, L. Struder, G. Hauser, H. Gorke, J. Ullrich, S. Herrmann, G. Schaller, F. Schopper, H. Soltau, K. U. Kuhnel, M. Messerschmidt, J. D. Bozek, S. P. Hau-Riege, M. Frank, C. Y. Hampton, R. G. Sierra, D. Starodub, G. J. Williams, J. Hajdu, N. Timneanu, M. M. Seibert, J. Andreasson, A. Rocker, O. Jonsson, M. Svenda, S. Stern, K. Nass, R. Andritschke, C. D. Schroter, F. Krasniqi, M. Bott, K. E. Schmidt, X. Y. Wang, I. Grotjohann, J. M. Holton, T. R. M. Barends, R. Neutze, S. Marchesini, R. Fromme, S. Schorb, D. Rupp, M. Adolph, T. Gorkhover, I. Andersson, H. Hirsemann, G. Potdevin, H. Graafsma, B. Nilsson and J. C. H. Spence, Nature, 2011, 470, U73–U81 CrossRef PubMed.
  58. D. P. DePonte, U. Weierstall, K. Schmidt, J. Warner, D. Starodub, J. C. H. Spence and R. B. Doak, J. Phys. D: Appl. Phys., 2008, 41, 195505 CrossRef.
  59. M. O. Wiedorn, D. Oberthur, R. Bean, R. Schubert, N. Werner, B. Abbey, M. Aepfelbacher, L. Adriano, A. Allahgholi, N. Al-Qudami, J. Andreasson, S. Aplin, S. Awel, K. Ayyer, S. Bajt, I. Barak, S. Bari, J. Bielecki, S. Botha, D. Boukhelef, W. Brehm, S. Brockhauser, I. Cheviakov, M. A. Coleman, F. Cruz-Mazo, C. Danilevski, C. Darmanin, R. B. Doak, M. Domaracky, K. Dorner, Y. Du, H. Fangohr, H. Fleckenstein, M. Frank, P. Fromme, A. M. Ganan-Calvo, Y. Gevorkov, K. Giewekemeyer, H. M. Ginn, H. Graafsma, R. Graceffa, D. Greiffenberg, L. Gumprecht, P. Gottlicher, J. Hajdu, S. Hauf, M. Heymann, S. Holmes, D. A. Horke, M. S. Hunter, S. Imlau, A. Kaukher, Y. Kim, A. Klyuev, J. Knoska, B. Kobe, M. Kuhn, C. Kupitz, J. Kuper, J. M. Lahey-Rudolph, T. Laurus, K. Le Cong, R. Letrun, P. L. Xavier, L. Maia, F. Maia, V. Mariani, M. Messerschmidt, M. Metz, D. Mezza, T. Michelat, G. Mills, D. C. F. Monteiro, A. Morgan, K. Muhlig, A. Munke, A. Munnich, J. Nette, K. A. Nugent, T. Nuguid, A. M. Orville, S. Pandey, G. Pena, P. Villanueva-Perez, J. Poehlsen, G. Previtali, L. Redecke, W. M. Riekehr, H. Rohde, A. Round, T. Safenreiter, I. Sarrou, T. Sato, M. Schmidt, B. Schmitt, R. Schonherr, J. Schulz, J. A. Sellberg, M. M. Seibert, C. Seuring, M. L. Shelby, R. L. Shoeman, M. Sikorski, A. Silenzi, C. A. Stan, X. T. Shi, S. Stern, J. Sztuk-Dambietz, J. Szuba, A. Tolstikova, M. Trebbin, U. Trunk, P. Vagovic, T. Ve, B. Weinhausen, T. A. White, K. Wrona, C. Xu, O. Yefanov, N. Zatsepin, J. G. Zhang, M. Perbandt, A. P. Mancuso, C. Betzel, H. Chapman and A. Barty, Nat. Commun., 2018, 9, 11 CrossRef PubMed.
  60. U. Weierstall, D. James, C. Wang, T. A. White, D. Wang, W. Liu, J. C. H. Spence, R. Bruce Doak, G. Nelson, P. Fromme, R. Fromme, I. Grotjohann, C. Kupitz, N. A. Zatsepin, H. Liu, S. Basu, D. Wacker, G. Won Han, V. Katritch, S. Boutet, M. Messerschmidt, G. J. Williams, J. E. Koglin, M. Marvin Seibert, M. Klinker, C. Gati, R. L. Shoeman, A. Barty, H. N. Chapman, R. A. Kirian, K. R. Beyerlein, R. C. Stevens, D. Li, S. T. A. Shah, N. Howe, M. Caffrey and V. Cherezov, Nat. Commun., 2014, 5, 3309 CrossRef.
  61. K. R. Beyerlein, H. O. Jönsson, R. Alonso-Mori, A. Aquila, S. Bajt, A. Barty, R. Bean, J. E. Koglin, M. Messerschmidt, D. Ragazzon, D. Sokaras, G. J. Williams, S. Hau-Riege, S. Boutet, H. N. Chapman, N. Tîmneanu and C. Caleman, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 5652–5657 CrossRef CAS.
  62. J. A. Sellberg, C. Huang, T. A. McQueen, N. D. Loh, H. Laksmono, D. Schlesinger, R. G. Sierra, D. Nordlund, C. Y. Hampton, D. Starodub, D. P. DePonte, M. Beye, C. Chen, A. V. Martin, A. Barty, K. T. Wikfeldt, T. M. Weiss, C. Caronna, J. Feldkamp, L. B. Skinner, M. M. Seibert, M. Messerschmidt, G. J. Williams, S. Boutet, L. G. M. Pettersson, M. J. Bogan and A. Nilsson, Nature, 2014, 510, 381–384 CrossRef CAS.
  63. H. Laksmono, T. A. McQueen, J. A. Sellberg, N. D. Loh, C. Huang, D. Schlesinger, R. G. Sierra, C. Y. Hampton, D. Nordlund, M. Beye, A. V. Martin, A. Barty, M. M. Seibert, M. Messerschmidt, G. J. Williams, S. Boutet, K. Amann-Winkel, T. Loerting, L. G. M. Pettersson, M. J. Bogan and A. Nilsson, J. Phys. Chem. Lett., 2015, 6, 2826–2832 CrossRef CAS PubMed.
  64. E. A. Schriber, D. W. Paley, R. Bolotovsky, D. J. Rosenberg, R. G. Sierra, A. Aquila, D. Mendez, F. Poitevin, J. P. Blaschke, A. Bhowmick, R. P. Kelly, M. Hunter, B. Hayes, D. C. Popple, M. Yeung, C. Pareja-Rivera, S. Lisova, K. Tono, M. Sugahara, S. Owada, T. Kuykendall, K. Yao, P. J. Schuck, D. Solis-Ibarra, N. K. Sauter, A. S. Brewster and J. N. Hohman, Nature, 2022, 601, 360–365 CrossRef CAS.
  65. F. Lehmkühler, F. Dallari, A. Jain, M. Sikorski, J. Möller, L. Frenzel, I. Lokteva, G. Mills, M. Walther, H. Sinn, F. Schulz, M. Dartsch, V. Markmann, R. Bean, Y. Kim, P. Vagovic, A. Madsen, P. Mancuso Adrian and G. Grübel, Proc. Natl. Acad. Sci. U. S. A., 2020, 117, 24110–24116 CrossRef.
  66. A. Aquila, M. S. Hunter, R. B. Doak, R. A. Kirian, P. Fromme, T. A. White, J. Andreasson, D. Arnlund, S. Bajt, T. R. M. Barends, M. Barthelmess, M. J. Bogan, C. Bostedt, H. Bottin, J. D. Bozek, C. Caleman, N. Coppola, J. Davidsson, D. P. DePonte, V. Elser, S. W. Epp, B. Erk, H. Fleckenstein, L. Foucar, M. Frank, R. Fromme, H. Graafsma, I. Grotjohann, L. Gumprecht, J. Hajdu, C. Y. Hampton, A. Hartmann, R. Hartmann, S. Hau-Riege, G. Hauser, H. Hirsemann, P. Holl, J. M. Holton, A. Hömke, L. Johansson, N. Kimmel, S. Kassemeyer, F. Krasniqi, K.-U. Kühnel, M. Liang, L. Lomb, E. Malmerberg, S. Marchesini, A. V. Martin, F. R. N. C. Maia, M. Messerschmidt, K. Nass, C. Reich, R. Neutze, D. Rolles, B. Rudek, A. Rudenko, I. Schlichting, C. Schmidt, K. E. Schmidt, J. Schulz, M. M. Seibert, R. L. Shoeman, R. Sierra, H. Soltau, D. Starodub, F. Stellato, S. Stern, L. Strüder, N. Timneanu, J. Ullrich, X. Wang, G. J. Williams, G. Weidenspointner, U. Weierstall, C. Wunderer, A. Barty, J. C. H. Spence and H. N. Chapman, Opt. Express, 2012, 20, 2706–2716 CrossRef CAS PubMed.
  67. V. Panneels, W. T. Wu, C. J. Tsai, P. Nogly, J. Rheinberger, K. Jaeger, G. Cicchetti, C. Gati, L. M. Kick, L. Sala, G. Capitani, C. Milne, C. Padeste, B. Pedrini, X. D. Li, J. Standfuss, R. Abela and G. Schertler, Struct. Dyn., 2015, 2, 8 CrossRef.
  68. M. Schmidt, Adv. Condens. Matter Phys., 2013, 10,  DOI:10.1155/2013/167276.
  69. G. D. Calvey, A. M. Katz and L. Pollack, Anal. Chem., 2019, 91, 7139–7144 CrossRef CAS PubMed.
  70. J. R. Stagno, Y. Liu, Y. R. Bhandari, C. E. Conrad, S. Panja, M. Swain, L. Fan, G. Nelson, C. Li, D. R. Wendel, T. A. White, J. D. Coe, M. O. Wiedorn, J. Knoska, D. Oberthuer, R. A. Tuckey, P. Yu, M. Dyba, S. G. Tarasov, U. Weierstall, T. D. Grant, C. D. Schwieters, J. Zhang, A. R. Ferre-D'Amare, P. Fromme, D. E. Draper, M. Liang, M. S. Hunter, S. Boutet, K. Tan, X. Zuo, X. Ji, A. Barty, N. A. Zatsepin, H. N. Chapman, J. C. H. Spence, S. A. Woodson and Y. X. Wang, Nature, 2017, 541, 242–246 CrossRef CAS.
  71. J. L. Olmos, S. Pandey, J. M. Martin-Garcia, G. Calvey, A. Katz, J. Knoska, C. Kupitz, M. S. Hunter, M. Liang, D. Oberthuer, O. Yefanov, M. Wiedorn, M. Heyman, M. Holl, K. Pande, A. Barty, M. D. Miller, S. Stern, S. Roy-Chowdhury, J. Coe, N. Nagaratnam, J. Zook, J. Verburgt, T. Norwood, I. Poudyal, D. Xu, J. Koglin, M. H. Seaberg, Y. Zhao, S. Bajt, T. Grant, V. Mariani, G. Nelson, G. Subramanian, E. Bae, R. Fromme, R. Fung, P. Schwander, M. Frank, T. A. White, U. Weierstall, N. Zatsepin, J. Spence, P. Fromme, H. N. Chapman, L. Pollack, L. Tremblay, A. Ourmazd, G. N. Phillips and M. Schmidt, BMC Biol., 2018, 16, 59 CrossRef.
  72. M. Dasgupta, D. Budday, S. H. P. de Oliveira, P. Madzelan, D. Marchany-Rivera, J. Seravalli, B. Hayes, R. G. Sierra, S. Boutet, M. S. Hunter, R. Alonso-Mori, A. Batyuk, J. Wierman, A. Lyubimov, A. S. Brewster, N. K. Sauter, G. A. Applegate, V. K. Tiwari, D. B. Berkowitz, M. C. Thompson, A. E. Cohen, J. S. Fraser, M. E. Wall, H. van den Bedem and M. A. Wilson, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 25634 CrossRef CAS PubMed.
  73. K. R. Beyerlein, D. Dierksmeyer, V. Mariani, M. Kuhn, I. Sarrou, A. Ottaviano, S. Awel, J. Knoska, S. Fuglerud, O. Jonsson, S. Stern, M. O. Wiedorn, O. Yefanov, L. Adriano, R. Bean, A. Burkhardt, P. Fischer, M. Heymann, D. A. Horke, K. E. J. Jungnickel, E. Kovaleva, O. Lorbeer, M. Metz, J. Meyer, A. Morgan, K. Pande, S. Panneerselvam, C. Seuring, A. Tolstikova, J. Lieske, S. Aplin, M. Roessle, T. A. White, H. N. Chapman, A. Meents and D. Oberthuer, IUCrJ, 2017, 4, 769–777 CrossRef CAS PubMed.
  74. R. Dimper, H. Reichert, P. Raimondi, L. Sánchez Ortiz, F. Sette and J. Susini, ESRF Upgrade Programme Phase II (2015-2022) - Technical Design Study (“The Orange Book”), The European Synchrotron Radiation Facility Search PubMed.
  75. A. Meents, M. O. Wiedorn, V. Srajer, R. Henning, I. Sarrou, J. Bergtholdt, M. Barthelmess, P. Y. A. Reinke, D. Dierksmeyer, A. Tolstikova, S. Schaible, M. Messerschmidt, C. M. Ogata, D. J. Kissick, M. H. Taft, D. J. Manstein, J. Lieske, D. Oberthuer, R. F. Fischetti and H. N. Chapman, Nat. Commun., 2017, 8, 12 CrossRef PubMed.
  76. P. Nogly, D. James, D. J. Wang, T. A. White, N. Zatsepin, A. Shilova, G. Nelson, H. G. Liu, L. Johansson, M. Heymann, K. Jaeger, M. Metz, C. Wickstrand, W. T. Wu, P. Bath, P. Berntsen, D. Oberthuer, V. Panneels, V. Cherezov, H. Chapman, G. Schertler, R. Neutze, J. Spence, I. Moraes, M. Burghammer, J. Standfuss and U. Weierstall, IUCrJ, 2015, 2, 168–176 CrossRef CAS PubMed.
  77. O. Saldanha, M. E. Brennich, M. Burghammer, H. Herrmann and S. Köster, Biomicrofluidics, 2016, 10, 024108 CrossRef.
  78. S. E. Wolf, J. Leiterer, M. Kappl, F. Emmerling and W. Tremel, J. Am. Chem. Soc., 2008, 130, 12342–12347 CrossRef CAS PubMed.
  79. P. Sonderby, C. Soderberg, C. G. Frankaer, G. Peters, J. T. Bukrinski, A. Labrador, T. S. Plivelic and P. Harris, J. Synchrotron Radiat., 2020, 27, 396–404 CrossRef.
  80. A. M. Seddon, S. J. Richardson, K. Rastogi, T. S. Plivelic, A. M. Squires and C. Pfrang, J. Phys. Chem. Lett., 2016, 7, 1341–1345 CrossRef CAS.
  81. A. Milsom, A. M. Squires, J. A. Boswell, N. J. Terrill, A. D. Ward and C. Pfrang, Atmos. Chem. Phys., 2021, 21, 15003–15021 CrossRef CAS.
  82. A. J. deMello, Nature, 2006, 442, 394–402 CrossRef CAS.
  83. Y. Liu and X. Jiang, Lab Chip, 2017, 17, 3960–3978 RSC.
  84. A. B. Theberge, F. Courtois, Y. Schaerli, M. Fischlechner, C. Abell, F. Hollfelder and W. T. S. Huck, Angew. Chem., Int. Ed., 2010, 49, 5846–5868 CrossRef CAS.
  85. S. Köster and T. Pfohl, Mod. Phys. Lett. B, 2012, 26, 1230018 CrossRef.
  86. H. Song, J. D. Tice and R. F. Ismagilov, Angew. Chem., Int. Ed., 2003, 42, 768–772 CrossRef CAS PubMed.
  87. B. M. Zwickl, W. E. Shanks, A. M. Jayich, C. Yang, A. C. Bleszynski Jayich, J. D. Thompson and J. G. E. Harris, Appl. Phys. Lett., 2008, 92, 103125 CrossRef.
  88. B. Weinhausen and S. Köster, Lab Chip, 2013, 13, 212–215 RSC.
  89. C. Gosse, S. Stanescu, J. Frederick, S. Lefrançois, A. Vecchiola, M. Moskura, S. Swaraj, R. Belkhou, B. Watts, P. Haltebourg, C. Blot, J. Daillant, P. Guenoun and C. Chevallard, Lab Chip, 2020, 20, 3213–3229 RSC.
  90. K. Dhouib, C. K. Malek, W. Pfleging, B. Gauthier-Manuel, R. Duffait, G. Thuillier, R. Ferrigno, L. Jacquamet, J. Ohana, J. L. Ferrer, A. Theobald-Dietrich, R. Giege, B. Lorber and C. Sauter, Lab Chip, 2009, 9, 1412–1421 RSC.
  91. M. E. Brennich, J. F. Nolting, C. Dammann, B. Noding, S. Bauch, H. Herrmann, T. Pfohl and S. Köster, Lab Chip, 2011, 11, 708–716 RSC.
  92. T. Lange, S. Charton, T. Bizien, F. Testard and F. Malloggi, Lab Chip, 2020, 20, 2990–3000 RSC.
  93. R. Barrett, M. Faucon, J. Lopez, G. Cristobal, F. Destremaut, A. Dodge, P. Guillot, P. Laval, C. Masselon and J. B. Salmon, Lab Chip, 2006, 6, 494–499 RSC.
  94. S. Sui, Y. X. Wang, K. W. Kolewe, V. Srajer, R. Henning, J. D. Schiffman, C. Dimitrakopoulos and S. L. Perry, Lab Chip, 2016, 16, 3082–3096 RSC.
  95. Z. Ren, M. Ayhan, S. Bandara, K. Bowatte, I. Kumarapperuma, S. Gunawardana, H. Shin, C. Wang, X. T. Zeng and X. J. Yang, Lab Chip, 2018, 18, 2246–2256 RSC.
  96. T. M. Squires and S. R. Quake, Rev. Mod. Phys., 2005, 77, 977–1026 CrossRef CAS.
  97. H. A. Stone, A. D. Stroock and A. Ajdari, Annu. Rev. Fluid Mech., 2004, 36, 381–411 CrossRef.
  98. Z. Lu, J. McMahon, H. Mohamed, D. Barnard, T. R. Shaikh, C. A. Mannella, T. Wagenknecht and T.-M. Lu, Sens. Actuators, B, 2010, 144, 301–309 CrossRef CAS.
  99. A. D. Stroock, S. K. W. Dertinger, A. Ajdari, I. Mezić, H. A. Stone and G. M. Whitesides, Science, 2002, 295, 647–651 CrossRef CAS.
  100. X. Casadevall i Solvas and A. deMello, Chem. Commun., 2011, 47, 1936–1942 RSC.
  101. O. Saldanha, R. Graceffa, C. Hemonnot, C. Ranke, G. Brehm, M. Liebi, B. Marmiroli, B. Weinhausen, M. Burghammer and S. Köster, ChemPhysChem, 2017, 18, 1220 CrossRef CAS PubMed.
  102. M. A. Levenstein, Y.-Y. Kim, L. Hunter, C. Anduix-Canto, C. González Niño, S. J. Day, S. Li, W. J. Marchant, P. A. Lee, C. C. Tang, M. Burghammer, F. C. Meldrum and N. Kapur, Lab Chip, 2020, 20, 2954–2964 RSC.
  103. B. Wunderlich, D. Nettels and B. Schuler, Lab Chip, 2014, 14, 219–228 RSC.
  104. B. D. Cullity and S. R. Stock, Elements of X-Ray Diffraction, Prentice Hall, New Jersey, 3rd edn, 2001 Search PubMed.
  105. P. Lindner and T. Zemb, Neutrons, X-rays and Light: Scattering Methods Applied to Soft Condensed Matter, North Holland, 2002 Search PubMed.
  106. A. Merlin, J. Angly, L. Daubersies, C. Madeira, S. Schoder, J. Leng and J. B. Salmon, Eur. Phys. J. E: Soft Matter Biol. Phys., 2011, 34, 1–7 CrossRef.
  107. L. C. McKenzie, P. M. Haben, S. D. Kevan and J. E. Hutchison, J. Phys. Chem. C, 2010, 114, 22055–22063 CrossRef CAS.
  108. M. Takesue, T. Tomura, M. Yamada, K. Hata, S. Kuwamoto and T. Yonezawa, J. Am. Chem. Soc., 2011, 133, 14164–14167 CrossRef CAS.
  109. R. Stehle, G. Goerigk, D. Wallacher, M. Ballauff and S. Seiffert, Lab Chip, 2013, 13, 1529–1537 RSC.
  110. I. Rodríguez-Ruiz, S. Charton, D. Radajewski, T. Bizien and S. Teychené, CrystEngComm, 2018, 20, 3302–3307 RSC.
  111. A. Y. Fong, L. Pellouchoud, M. Davidson, R. C. Walroth, C. Church, E. Tcareva, L. Wu, K. Peterson, B. Meredig and C. J. Tassone, J. Chem. Phys., 2021, 154, 224201 CrossRef CAS.
  112. K. Younes, M. Poli, P. Muhunthan, I. Rajkovic, S. Ermon, T. M. Weiss and M. Ihme, Nucl. Instrum. Methods Phys. Res., Sect. A, 2023, 1057, 168719 CrossRef CAS.
  113. J. F. Moulin, S. V. Roth and P. Müller-Buschbaum, Rev. Sci. Instrum., 2008, 79, 015109 CrossRef.
  114. E. Metwalli, J. F. Moulin, J. Perlich, W. Wang, A. Diethert, S. V. Roth and P. Müller-Buschbaum, Langmuir, 2009, 25, 11815–11821 CrossRef CAS PubMed.
  115. J. Kehres, T. Pedersen, F. Masini, J. W. Andreasen, M. M. Nielsen, A. Diaz, J. H. Nielsen, O. Hansen and I. Chorkendorff, J. Synchrotron Radiat., 2016, 23, 455–463 CrossRef CAS PubMed.
  116. H. G. Alison, R. J. Davey, J. Garside, M. J. Quayle, G. J. T. Tiddy, D. T. Clarke and G. R. Jones, Phys. Chem. Chem. Phys., 2003, 5, 4998–5000 RSC.
  117. T. Chen, A. Neville, K. Sorbie and Z. Zhong, Faraday Discuss., 2007, 136, 355–365 RSC.
  118. T. Chen, A. Neville, K. Sorbie and Z. Zhong, Chem. Eng. Sci., 2009, 64, 912–918 CrossRef CAS.
  119. E. Mavredaki, A. Neville and K. Sorbie, Appl. Surf. Sci., 2011, 257, 4264–4271 CrossRef CAS.
  120. D. Burkle, R. De Motte, W. Taleb, A. Kleppe, T. Comyn, S. M. Vargas, A. Neville and R. Barker, Rev. Sci. Instrum., 2016, 87, 7 CrossRef.
  121. T. Beuvier, E. A. C. Panduro, P. Kwasniewski, S. Marre, C. Lecoutre, Y. Garrabos, C. Aymonier, B. Calvignac and A. Gibaud, Lab Chip, 2015, 15, 2002–2008 RSC.
  122. M. A. Levenstein, C. Anduix-Canto, Y.-Y. Kim, M. A. Holden, C. González Niño, D. C. Green, S. E. Foster, A. N. Kulak, L. Govada, N. E. Chayen, S. J. Day, C. C. Tang, B. Weinhausen, M. Burghammer, N. Kapur and F. C. Meldrum, Adv. Funct. Mater., 2019, 29, 1808172 CrossRef.
  123. M. A. Levenstein, L. E. Wayment, C. D. Scott, R. A. Lunt, P.-B. Flandrin, S. Day, C. Tang, C. C. Wilson, F. C. Meldrum, N. Kapur and K. Robertson, Anal. Chem., 2020, 92, 7754–7761 CrossRef CAS PubMed.
  124. D. Radajewski, L. Hunter, X. He, O. Nahi, J. M. Galloway and F. C. Meldrum, Lab Chip, 2021, 21, 4498–4506 RSC.
  125. P. J. E. M. van der Linden, A. M. Popov and D. Pontoni, Lab Chip, 2020, 20, 4128–4140 RSC.
  126. B. Fleury, M.-A. Neouze, J.-M. Guigner, N. Menguy, O. Spalla, T. Gacoin and D. Carriere, ACS Nano, 2014, 8, 2602–2608 CrossRef CAS.
  127. M. O. Besenhard, A. P. LaGrow, A. Hodzic, M. Kriechbaum, L. Panariello, G. Bais, K. Loizou, S. Damilos, M. Margarida Cruz, N. T. K. Thanh and A. Gavriilidis, Chem. Eng. J., 2020, 399, 125740 CrossRef CAS.
  128. M. Durelle, F. Gobeaux, T. K. Truong, S. Charton and D. Carriere, Cryst. Growth Des., 2023, 23, 5631–5640 CrossRef CAS.
  129. M. W. Terban, D. Banerjee, S. Ghose, B. Medasani, A. Shukla, B. A. Legg, Y. Zhou, Z. Zhu, M. L. Sushko, J. J. De Yoreo, J. Liu, P. K. Thallapally and S. J. L. Billinge, Nanoscale, 2018, 10, 4291–4300 RSC.
  130. M. L. Beauvais, P. J. Chupas, D. O'Nolan, J. B. Parise and K. W. Chapman, ACS Mater. Lett., 2021, 3, 698–703 CrossRef CAS.
  131. M. L. Beauvais, B. A. Sanchez Monserrate, T. Feng, R. Chen, P. Liu, P. J. Chupas and K. W. Chapman, J. Appl. Crystallogr., 2022, 55, 258–264 CrossRef CAS.
  132. M. J. Young, N. M. Bedford, N. Jiang, D. Lin and L. Dai, J. Synchrotron Radiat., 2017, 24, 787–795 CrossRef CAS PubMed.
  133. G. Kwon, Y.-H. Cho, K.-B. Kim, J. D. Emery, I. S. Kim, X. Zhang, A. B. F. Martinson and D. M. Tiede, J. Synchrotron Radiat., 2019, 26, 1600–1611 CrossRef CAS PubMed.
  134. L. Pollack, M. W. Tate, N. C. Darnton, J. B. Knight, S. M. Gruner, W. A. Eaton and R. H. Austin, Proc. Natl. Acad. Sci. U. S. A., 1999, 96, 10115–10117 CrossRef CAS.
  135. L. Pollack, M. W. Tate, A. C. Finnefrock, C. Kalidas, S. Trotter, N. C. Darnton, L. Lurio, R. H. Austin, C. A. Batt, S. M. Gruner and S. G. J. Mochrie, Phys. Rev. Lett., 2001, 86, 4962–4965 CrossRef CAS PubMed.
  136. M. Jiang and R. D. Braatz, CrystEngComm, 2019, 21, 3534–3551 RSC.
  137. M. O. Besenhard, S. Pal, G. Gkogkos and A. Gavriilidis, React. Chem. Eng., 2023, 8, 955–977 RSC.
  138. K. Robertson, P. B. Flandrin, A. R. Klapwijk and C. C. Wilson, Cryst. Growth Des., 2016, 16, 4759–4764 CrossRef CAS.
  139. V. Körstgens, M. Philipp, D. Magerl, M. A. Niedermeier, G. Santoro, S. V. Roth and P. Müller-Buschbaum, RSC Adv., 2014, 4, 1476–1479 RSC.
  140. J. Han, F. Testard, F. Malloggi, C. Pierre-Eugene, N. Menguy and O. Spalla, Langmuir, 2012, 28, 15966–15974 CrossRef PubMed.
  141. R. K. Ramamoorthy, E. Yildirim, E. Barba, P. Roblin, J. A. Vargas, L.-M. Lacroix, I. Rodriguez-Ruiz, P. Decorse, V. Petkov, S. Teychené and G. Viau, Nanoscale, 2020, 12, 16173–16188 RSC.
  142. B. He, L. K. Macreadie, J. Gardiner, S. G. Telfer and M. R. Hill, ACS Appl. Mater. Interfaces, 2021, 13, 54284–54293 CrossRef CAS.
  143. X. Chen, J. Wang, R. Pan, S. Roth and S. Förster, J. Phys. Chem. C, 2021, 125, 1087–1095 CrossRef CAS.
  144. S. J. L. Billinge and M. G. Kanatzidis, Chem. Commun., 2004, 749–760,  10.1039/b309577k.
  145. B. Winter, Nucl. Instrum. Methods Phys. Res., Sect. A, 2009, 601, 139–150 CrossRef CAS.
  146. A. Iglesias-Juez, G. L. Chiarello, G. S. Patience and M. O. Guerrero-Pérez, Can. J. Chem. Eng., 2022, 100, 3–22 CrossRef CAS.
  147. O. Proux, E. Lahera, W. Del Net, I. Kieffer, M. Rovezzi, D. Testemale, M. Irar, S. Thomas, A. Aguilar-Tapia, E. F. Bazarkina, A. Prat, M. Tella, M. Auffan, J. Rose and J.-L. Hazemann, J. Environ. Qual., 2017, 46, 1146–1157 CrossRef CAS PubMed.
  148. M. A. Newton, A. J. Dent and J. Evans, Chem. Soc. Rev., 2002, 31, 83–95 RSC.
  149. K. Amemiya, K. Sakata and M. Suzuki-Sakamaki, Rev. Sci. Instrum., 2020, 91, 093104 CrossRef CAS.
  150. B. Detlefs, S. Graziano and P. Glatzel, Anal. Chem., 2023, 95, 8758–8762 CrossRef CAS.
  151. M. C. Ringo, M. S. Huhta, G. Shea-McCarthy, J. E. Penner-Hahn and C. E. Evans, Nucl. Instrum. Methods Phys. Res., Sect. B, 1999, 149, 177–181 CrossRef CAS.
  152. S. E. Mann, M. C. Ringo, G. Shea-McCarthy, J. Penner-Hahn and C. E. Evans, Anal. Chem., 2000, 72, 1754–1758 CrossRef CAS PubMed.
  153. M. M. Hoffmann, J. G. Darab, S. M. Heald, C. R. Yonker and J. L. Fulton, Chem. Geol., 2000, 167, 89–103 CrossRef CAS.
  154. M. M. Hoffmann, J. G. Darab and J. L. Fulton, Rev. Sci. Instrum., 2001, 105, 6876–6885 CAS.
  155. G. Sankar, E. Cao and A. Gavriilidis, Catal. Today, 2007, 125, 24–28 CrossRef CAS.
  156. E. M. Chan, M. A. Marcus, S. Fakra, M. ElNaggar, R. A. Mathies and A. P. Alivisatos, J. Phys. Chem. A, 2007, 111, 12210–12215 CrossRef CAS.
  157. G. Hofmann, G. Tofighi, G. Rinke, S. Baier, A. Ewinger, A. Urban, A. Wenka, S. Heideker, A. Jahn, R. Dittmeyer and J. D. Grunwaldt, J. Phys.: Conf. Ser., 2016, 712, 012072 CrossRef.
  158. G. Tofighi, H. Lichtenberg, J. Pesek, T. L. Sheppard, W. Wang, L. Schöttner, G. Rinke, R. Dittmeyer and J.-D. Grunwaldt, React. Chem. Eng., 2017, 2, 876–884 RSC.
  159. P. Micheal Raj, L. Barbe, M. Andersson, M. De Albuquerque Moreira, D. Haase, J. Wootton, S. Nehzati, A. E. Terry, R. J. Friel, M. Tenje and K. G. V. Sigfridsson Clauss, RSC Adv., 2021, 11, 29859–29869 RSC.
  160. S. Britto, C. M. A. Parlett, S. Bartlett, J. D. Elliott, K. Ignatyev and S. L. M. Schroeder, J. Phys. Chem. C, 2023, 127, 8631–8639 CrossRef CAS.
  161. S. Zinoveva, R. De Silva, R. D. Louis, P. Datta, C. S. S. R. Kumar, J. Goettert and J. Hormes, Nucl. Instrum. Methods Phys. Res., Sect. A, 2007, 582, 239–241 CrossRef CAS.
  162. H. Oyanagi, Z. H. Sun, Y. Jiang, M. Uehara, H. Nakamura, K. Yamashita, L. Zhang, C. Lee, A. Fukano and H. Maeda, J. Synchrotron Radiat., 2011, 18, 272–279 CrossRef CAS.
  163. J. Monnier, L. Legrand, L. Bellot-Gurlet, E. Foy, S. Reguer, E. Rocca, P. Dillmann, D. Neff, F. Mirambet, S. Perrin and I. Guillot, J. Nucl. Mater., 2008, 379, 105–111 CrossRef CAS.
  164. J. Monnier, S. Réguer, E. Foy, D. Testemale, F. Mirambet, M. Saheb, P. Dillmann and I. Guillot, Corros. Sci., 2014, 78, 293–303 CrossRef CAS.
  165. K. S. Krishna, C. V. Navin, S. Biswas, V. Singh, K. Ham, G. L. Bovenkamp, C. S. Theegala, J. T. Miller, J. J. Spivey and C. S. S. R. Kumar, J. Am. Chem. Soc., 2013, 135, 5450–5456 CrossRef PubMed.
  166. A. V. Dobrovolskaya, S. V. Chapek, O. A. Usoltsev, E. Naranov, D. N. Gorbunov, A. L. Trigub, A. L. Maximov, A. V. Soldatov and A. L. Bugaev, J. Phys. Chem. C, 2023, 127, 20727–20733 CrossRef CAS.
  167. J. L. Fulton, Y. Chen, S. M. Heald and M. Balasubramanian, Rev. Sci. Instrum., 2004, 75, 5228–5231 CrossRef CAS.
  168. O. Fuchs, F. Maier, L. Weinhardt, M. Weigand, M. Blum, M. Zharnikov, J. Denlinger, M. Grunze, C. Heske and E. Umbach, Nucl. Instrum. Methods Phys. Res., Sect. A, 2008, 585, 172–177 CrossRef CAS.
  169. M. Nagasaka, T. Hatsui, T. Horigome, Y. Hamamura and N. Kosugi, J. Electron Spectrosc. Relat. Phenom., 2010, 177, 130–134 CrossRef CAS.
  170. S. Schreck, G. Gavrila, C. Weniger and P. Wernet, Rev. Sci. Instrum., 2011, 82, 103101 CrossRef.
  171. J. Probst, C. N. Borca, M. A. Newton, J. van Bokhoven, T. Huthwelker, S. Stavrakis and A. deMello, ACS Meas. Sci. Au, 2021, 1, 27–34 CrossRef CAS.
  172. J. Brenker, K. Henzler, C. N. Borca, T. Huthwelker and T. Alan, Lab Chip, 2022, 22, 1214–1230 RSC.
  173. G. L. Chiarello, M. Nachtegaal, V. Marchionni, L. Quaroni and D. Ferri, Rev. Sci. Instrum., 2014, 85, 074102 CrossRef.
  174. E. K. Dann, E. K. Gibson, C. R. A. Catlow, V. Celorrio, P. Collier, T. Eralp, M. Amboage, C. Hardacre, C. Stere, A. Kroner, A. Raj, S. Rogers, A. Goguet and P. P. Wells, J. Catal., 2019, 373, 201–208 CrossRef CAS.
  175. B. Venezia, E. Cao, S. K. Matam, C. Waldron, G. Cibin, E. K. Gibson, S. Golunski, P. P. Wells, I. Silverwood, C. R. A. Catlow, G. Sankar and A. Gavriilidis, Catal. Sci. Technol., 2020, 10, 7842–7856 RSC.
  176. D. A. Huyke, A. Ramachandran, O. Ramirez-Neri, J. A. Guerrero-Cruz, L. B. Gee, A. Braun, D. Sokaras, B. Garcia-Estrada, E. I. Solomon, B. Hedman, M. U. Delgado-Jaime, D. P. DePonte, T. Kroll and J. G. Santiago, J. Synchrotron Radiat., 2021, 28, 1100–1113 CrossRef CAS.
  177. A. M. Karim, N. Al Hasan, S. Ivanov, S. Siefert, R. T. Kelly, N. G. Hallfors, A. Benavidez, L. Kovarik, A. Jenkins, R. E. Winans and A. K. Datye, J. Phys. Chem. C, 2015, 119, 13257–13267 CrossRef.
  178. T. Binninger, E. Fabbri, A. Patru, M. Garganourakis, J. Han, D. F. Abbott, O. Sereda, R. Kötz, A. Menzel, M. Nachtegaal and T. J. Schmidt, J. Electrochem. Soc., 2016, 163, H906 CrossRef CAS.
  179. R. K. Ramamoorthy, E. Yildirim, I. Rodriguez-Ruiz, P. Roblin, L.-M. Lacroix, A. Diaz, R. Parmar, S. Teychené and G. Viau, Lab Chip, 2024, 24, 327–338 RSC.
  180. S. Busch, T. H. Jensen, Y. Chushkin and A. Fluerasu, Eur. Phys. J. E: Soft Matter Biol. Phys., 2008, 26, 55–62 CrossRef CAS.
  181. A. Fluerasu, A. Moussaid, P. Falus, H. Gleyzolle and A. Madsen, J. Synchrotron Radiat., 2008, 15, 378–384 CrossRef CAS PubMed.
  182. A. Fluerasu, P. Kwasniewski, C. Caronna, F. Destremaut, J.-B. Salmon and A. Madsen, New J. Phys., 2010, 12, 035023 CrossRef.
  183. R. Urbani, F. Westermeier, B. Banusch, M. Sprung and T. Pfohl, J. Synchrotron Radiat., 2016, 23, 1401–1408 CrossRef CAS PubMed.
  184. J. R. M. Lhermitte, M. C. Rogers, S. Manet and M. Sutton, Rev. Sci. Instrum., 2017, 88, 015112 CrossRef PubMed.
  185. H.-G. Steinrück, C. J. Takacs, H.-K. Kim, D. G. Mackanic, B. Holladay, C. Cao, S. Narayanan, E. M. Dufresne, Y. Chushkin, B. Ruta, F. Zontone, J. Will, O. Borodin, S. K. Sinha, V. Srinivasan and M. F. Toney, Energy Environ. Sci., 2020, 13, 4312–4321 RSC.
  186. P. Muhunthan, H. Li, G. Vignat, E. R. Toro, K. Younes, Y. Sun, D. Sokaras, T. Weiss, I. Rajkovic, T. Osaka, I. Inoue, S. Song, T. Sato, D. Zhu, J. L. Fulton and M. Ihme, Rev. Sci. Instrum., 2024, 95, 013901 CrossRef CAS PubMed.
  187. R. L. Leheny, Curr. Opin. Colloid Interface Sci., 2012, 17, 3–12 CrossRef CAS.
  188. O. Shpyrko, J. Synchrotron Radiat., 2014, 21, 1057–1064 CrossRef CAS PubMed.
  189. A. Sakdinawat and D. Attwood, Nat. Photonics, 2010, 4, 840–848 CrossRef CAS.
  190. F. Pfeiffer, Nat. Photonics, 2018, 12, 9–17 CrossRef CAS.
  191. M. Endrizzi, Nucl. Instrum. Methods Phys. Res., Sect. A, 2018, 878, 88–98 CrossRef CAS.
  192. P. J. Withers, C. Bouman, S. Carmignato, V. Cnudde, D. Grimaldi, C. K. Hagen, E. Maire, M. Manley, A. Du Plessis and S. R. Stock, Nat. Rev. Methods Primers, 2021, 1, 18 CrossRef CAS.
  193. U. Neuhausler, C. Jacobsen, D. Schulze, D. Stott and S. Abend, J. Synchrotron Radiat., 2000, 7, 110–112 CrossRef CAS PubMed.
  194. J. Rieger, J. Thieme and C. Schmidt, Langmuir, 2000, 16, 8300–8305 CrossRef CAS.
  195. D. Guay, J. Stewart-Ornstein, X. Zhang and A. P. Hitchcock, Anal. Chem., 2005, 77, 3479–3487 CrossRef CAS.
  196. I. J. Drake, T. C. N. Liu, M. Gilles, T. Tyliszczak, A. L. D. Kilcoyne, D. K. Shuh, R. A. Mathies and A. T. Bell, Rev. Sci. Instrum., 2004, 75, 3242–3247 CrossRef CAS.
  197. T. Huthwelker, V. Zelenay, M. Birrer, A. Krepelova, J. Raabe, G. Tzvetkov, M. G. C. Vernooij and M. Ammann, Rev. Sci. Instrum., 2010, 81, 113706 CrossRef CAS.
  198. V. Zelenay, M. Ammann, A. Křepelová, M. Birrer, G. Tzvetkov, M. G. C. Vernooij, J. Raabe and T. Huthwelker, J. Aerosol Sci., 2011, 42, 38–51 CrossRef CAS.
  199. S. T. Kelly, P. Nigge, S. Prakash, A. Laskin, B. Wang, T. Tyliszczak, S. R. Leone and M. K. Gilles, Rev. Sci. Instrum., 2013, 84, 073708 CrossRef.
  200. E. de Smit, I. Swart, J. F. Creemer, G. H. Hoveling, M. K. Gilles, T. Tyliszczak, P. J. Kooyman, H. W. Zandbergen, C. Morin, B. M. Weckhuysen and F. M. F. de Groot, Nature, 2008, 456, 222–225 CrossRef CAS.
  201. E. de Smit, I. Swart, J. F. Creemer, C. Karunakaran, D. Bertwistle, H. W. Zandbergen, F. M. F. de Groot and B. M. Weckhuysen, Angew. Chem., Int. Ed., 2009, 48, 3632–3636 CrossRef CAS.
  202. M. Yoo, Y.-S. Yu, H. Ha, S. Lee, J.-S. Choi, S. Oh, E. Kang, H. Choi, H. An, K.-S. Lee, J. Y. Park, R. Celestre, M. A. Marcus, K. Nowrouzi, D. Taube, D. A. Shapiro, W. Jung, C. Kim and H. Y. Kim, Energy Environ. Sci., 2020, 13, 1231–1239 RSC.
  203. Y. Sun and Y. Wang, Nano Lett., 2011, 11, 4386–4392 CrossRef CAS.
  204. J. Lim, Y. Li, D. H. Alsem, H. So, S. C. Lee, P. Bai, D. A. Cogswell, X. Liu, N. Jin, Y.-s. Yu, N. J. Salmon, D. A. Shapiro, M. Z. Bazant, T. Tyliszczak and W. C. Chueh, Science, 2016, 353, 566–571 CrossRef CAS PubMed.
  205. T. Mefford, K. Karki, D. H. Alsem, D. Shapiro, N. Salmon and W. C. Chueh, Microsc. Microanal., 2019, 25, 2094–2095 CrossRef.
  206. J. T. Mefford, A. R. Akbashev, M. Kang, C. L. Bentley, W. E. Gent, H. D. Deng, D. H. Alsem, Y.-S. Yu, N. J. Salmon, D. A. Shapiro, P. R. Unwin and W. C. Chueh, Nature, 2021, 593, 67–73 CrossRef CAS.
  207. T. Ohigashi, M. Nagasaka, T. Horigome, N. Kosugi, S. M. Rosendahl and A. P. Hitchcock, AIP Conf. Proc., 2016, 1741, 050002 CrossRef.
  208. V. Prabu, M. Obst, H. Hosseinkhannazer, M. Reynolds, S. Rosendahl, J. Wang and A. P. Hitchcock, Rev. Sci. Instrum., 2018, 89, 063702 CrossRef PubMed.
  209. S.-J. Lee and G.-B. Kim, J. Appl. Phys., 2003, 94, 3620–3623 CrossRef CAS.
  210. W.-K. Lee, K. Fezzaa and T. Uemura, J. Synchrotron Radiat., 2011, 18, 302–304 CrossRef.
  211. M. Prodanović, W. B. Lindquist and R. S. Seright, J. Colloid Interface Sci., 2006, 298, 282–297 CrossRef.
  212. C. Noiriel, B. Madé and P. Gouze, Water Resour. Res., 2007, 43 DOI:10.1029/2006WR005379.
  213. S. Berg, H. Ott, S. A. Klapp, A. Schwing, R. Neiteler, N. Brussee, A. Makurat, L. Leu, F. Enzmann, J.-O. Schwarz, M. Kersten, S. Irvine and M. Stampanoni, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 3755–3759 CrossRef CAS.
  214. R. T. Armstrong, H. Ott, A. Georgiadis, M. Rücker, A. Schwing and S. Berg, Water Resour. Res., 2014, 50, 9162–9176 CrossRef.
  215. S. Hasan, V. Niasar, N. K. Karadimitriou, J. R. A. Godinho, N. T. Vo, S. An, A. Rabbani and H. Steeb, Proc. Natl. Acad. Sci. U. S. A., 2020, 117, 23443–23449 CrossRef CAS.
  216. A. Piovesan, T. Van De Looverbosch, P. Verboven, C. Achille, C. Parra Cabrera, E. Boller, Y. Cheng, R. Ameloot and B. Nicolai, Lab Chip, 2020, 20, 2403–2411 RSC.
  217. K. J. Dobson, S. B. Coban, S. A. McDonald, J. N. Walsh, R. C. Atwood and P. J. Withers, Solid Earth, 2016, 7, 1059–1073 CrossRef.
  218. T. Bultreys, S. Ellman, C. M. Schlepütz, M. N. Boone, G. K. Pakkaner, S. Wang, M. Borji, S. Van Offenwert, N. Moazami Goudarzi, W. Goethals, C. W. Winardhi and V. Cnudde, Proc. Natl. Acad. Sci. U. S. A., 2024, 121, e2316723121 CrossRef CAS.
  219. J. Knoška, L. Adriano, S. Awel, K. R. Beyerlein, O. Yefanov, D. Oberthuer, G. E. Peña Murillo, N. Roth, I. Sarrou, P. Villanueva-Perez, M. O. Wiedorn, F. Wilde, S. Bajt, H. N. Chapman and M. Heymann, Nat. Commun., 2020, 11, 657 CrossRef.
  220. C. Noiriel, P. Gouze and B. Madé, J. Hydrol., 2013, 486, 211–223 CrossRef CAS.
  221. J. R. A. Godinho, K. M. Gerke, A. G. Stack and P. D. Lee, Sci. Rep., 2016, 6, 33086 CrossRef CAS PubMed.
  222. F. Fusseis, H. Steeb, X. Xiao, W.-l. Zhu, I. B. Butler, S. Elphick and U. Mader, J. Synchrotron Radiat., 2014, 21, 251–253 CrossRef CAS PubMed.
  223. S. Morais, C. Lecoutre, G. Philippot, G. Aubert, O. Nguyen, A. Cario, E. Vidal, Z. S. Campbell, Y. Garrabos, M. Azaroual, L. Helfen, D. Bernard and S. Marre, Processes, 2023, 11, 1981 CrossRef CAS.
  224. J. R. A. Godinho, L. Ma, Y. Chai, M. Storm and T. L. Burnett, Minerals, 2019, 9, 480 CrossRef CAS.
  225. C. Anduix-Canto, M. A. Levenstein, Y.-Y. Kim, J. R. A. Godinho, A. N. Kulak, C. G. Niño, P. J. Withers, J. P. Wright, N. Kapur, H. K. Christenson and F. C. Meldrum, Adv. Funct. Mater., 2021, 31, 2107312 CrossRef CAS.
  226. M. Nagasaka, H. Yuzawa, N. Takada, M. Aoyama, E. Rühl and N. Kosugi, J. Chem. Phys., 2019, 151, 114201 CrossRef.
  227. I. Chaussavoine, A. Beauvois, T. Mateo, R. Vasireddi, N. Douri, J. Priam, Y. Liatimi, S. Lefrancois, H. Tabuteau, M. Davranche, D. Vantelon, T. Bizien, L. M. G. Chavas and B. Lassalle-Kaiser, J. Synchrotron Radiat., 2020, 27, 230–237 CrossRef CAS PubMed.
  228. M. A. Chen and B. D. Kocar, J. Synchrotron Radiat., 2021, 28, 461–471 CrossRef CAS PubMed.
  229. I. T. Neckel, L. F. de Castro, F. Callefo, V. C. Teixeira, A. L. Gobbi, M. H. Piazzetta, R. A. G. de Oliveira, R. S. Lima, R. A. Vicente, D. Galante and H. C. N. Tolentino, Sci. Rep., 2021, 11, 23671 CrossRef CAS PubMed.
  230. Y. Matsumoto, Y. Takeo, S. Egawa, G. Yamaguchi, S. Yokomae, M. Takei, H. Yumoto, T. Koyama, H. Ohashi, K. Tono, M. Yabashi, H. Mimura and T. Kimura, Opt. Rev., 2022, 29, 7–12 CrossRef CAS.
  231. K. K. Goncz, P. Batson, D. Ciarlo, B. W. Loo Jr and S. S. Rothman, J. Microsc., 1992, 168, 101–110 CrossRef.
  232. J. Pine and J. R. Gilbert, Live Cell Specimens for X-Ray Microscopy, X-Ray Microscopy III: Proceedings of the Third International Conference, Berlin, Heidelberg, 1992 Search PubMed.
  233. F. M. F. de Groot, E. de Smit, M. M. van Schooneveld, L. R. Aramburo and B. M. Weckhuysen, ChemPhysChem, 2010, 11, 951–962 CrossRef CAS PubMed.
  234. B. Bozzini, M. K. Abyaneh, M. Amati, A. Gianoncelli, L. Gregoratti, B. Kaulich and M. Kiskinova, Chem. – Eur. J., 2012, 18, 10196–10210 CrossRef CAS.
  235. J. F. Creemer, S. Helveg, G. H. Hoveling, S. Ullmann, A. M. Molenbroek, P. M. Sarro and H. W. Zandbergen, Ultramicroscopy, 2008, 108, 993–998 CrossRef CAS.
  236. B. Bozzini, A. Gianoncelli, P. Bocchetta, S. Dal Zilio and G. Kourousias, Anal. Chem., 2014, 86, 664–670 CrossRef CAS PubMed.
  237. T. J. Heindel, J. Fluids Eng., 2011, 133, 074001 CrossRef.
  238. A. Snigirev, I. Snigireva, V. Kohn, S. Kuznetsov and I. Schelokov, Rev. Sci. Instrum., 1995, 66, 5486–5492 CrossRef CAS.
  239. C. Noiriel, P. Gouze and D. Bernard, Geophys. Res. Lett., 2004, 31, L24603 CrossRef.
  240. M. Di Michiel, J. M. Merino, D. Fernandez-Carreiras, T. Buslaps, V. Honkimäki, P. Falus, T. Martins and O. Svensson, Rev. Sci. Instrum., 2005, 76, 043702 CrossRef.
  241. R. Mokso, F. Marone, D. Haberthür, J. C. Schittny, G. Mikuljan, A. Isenegger and M. Stampanoni, AIP Conf. Proc., 2011, 1365, 38–41 CrossRef.
  242. A. Georgiadis, S. Berg, A. Makurat, G. Maitland and H. Ott, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2013, 88, 033002 CrossRef CAS PubMed.
  243. S. Youssef, H. Deschamps, J. Dautriat, E. Rosenberg, R. Oughanem, E. Maire and R. Mokso, presented in part at the International Symposium of the Society of Core Analysts, SCA2013-012, Napa Valley, California, USA, 16-19 September, 2013 Search PubMed.
  244. S. Schlüter, S. Berg, M. Rücker, R. T. Armstrong, H. J. Vogel, R. Hilfer and D. Wildenschild, Water Resour. Res., 2016, 52, 2194–2205 CrossRef.
  245. K. Singh, H. Menke, M. Andrew, Q. Lin, C. Rau, M. J. Blunt and B. Bijeljic, Sci. Rep., 2017, 7, 5192 CrossRef PubMed.
  246. A. Jahanbakhsh, K. L. Wlodarczyk, D. P. Hand, R. R. J. Maier and M. M. Maroto-Valer, Sensors, 2020, 20, 4030 CrossRef CAS PubMed.
  247. A. Aliseda and T. J. Heindel, Annu. Rev. Fluid Mech., 2021, 53, 543–567 CrossRef.
  248. L. Helfen, T. Baumbach, P. Mikulík, D. Kiel, P. Pernot, P. Cloetens and J. Baruchel, Appl. Phys. Lett., 2005, 86, 071915 CrossRef.
  249. C. Noiriel and F. Renard, C. R. Geosci., 2022, 354, 255–280 CrossRef.
  250. K. H. Cats and B. M. Weckhuysen, ChemCatChem, 2016, 8, 1531–1542 CrossRef CAS PubMed.
  251. H. Gesswein, P. Stuble, D. Weber, J. R. Binder and R. Monig, J. Appl. Crystallogr., 2022, 55, 503–514 CrossRef CAS PubMed.
  252. J. Lyngso and J. S. Pedersen, J. Appl. Crystallogr., 2021, 54, 295–305 CrossRef.
  253. P. Yuriy, V. Boris, J. Licai and K. Bonglea, Beam conditioning multilayer optics for laboratory x-ray sources, Advances in Laboratory-based X-Ray Sources, Optics, and Applications IV, 9590, San Diego, CA, 2015 Search PubMed.
  254. O. Taché, S. Rouziere, P. Joly, M. Amara, B. Fleury, A. Thill, P. Launois, O. Spalla and B. Abecassis, J. Appl. Crystallogr., 2016, 49, 1624–1631 CrossRef.
  255. P. Le Magueres, M. DelCampo, K. Saito, J. D. Ferrara, J. Wojciechowski, A. Jones, D. Kucharczyk and M. Meyer, Acta Crystallogr., Sect. A: Found. Adv., 2019, 75, a163 Search PubMed.
  256. D. J. Batey, F. Van Assche, S. Vanheule, M. N. Boone, A. J. Parnell, O. O. Mykhaylyk, C. Rau and S. Cipiccia, Phys. Rev. Lett., 2021, 126, 193902 CrossRef CAS PubMed.
  257. J. Polte, R. Erler, A. F. Thunemann, S. Sokolov, T. T. Ahner, K. Rademann, F. Emmerling and R. Kraehnert, ACS Nano, 2010, 4, 1076–1082 CrossRef CAS.
  258. J. Polte, X. Tuaev, M. Wuithschick, A. Fischer, A. F. Thuenemann, K. Rademann, R. Kraehnert and F. Emmerling, ACS Nano, 2012, 6, 5791–5802 CrossRef CAS.
  259. F. Kettemann, M. Wuithschick, G. Caputo, R. Kraehnert, N. Pinna, K. Rademann and J. Polte, CrystEngComm, 2015, 17, 1865–1870 RSC.
  260. X. Chen, J. Schröder, S. Hauschild, S. Rosenfeldt, M. Dulle and S. Förster, Langmuir, 2015, 31, 11678–11691 CrossRef CAS PubMed.
  261. M. Herbst, E. Hofmann and S. Förster, Langmuir, 2019, 35, 11702–11709 CrossRef CAS.
  262. J. Tillier, T. Binninger, M. Garganourakis, A. Patru, E. Fabbri, T. J. Schmidt and O. Sereda, J. Electrochem. Soc., 2016, 163, H913 CrossRef CAS.
  263. S. Bucciarelli, S. R. Midtgaard, M. Nors Pedersen, S. Skou, L. Arleth and B. Vestergaard, J. Appl. Crystallogr., 2018, 51, 1623–1632 CrossRef CAS PubMed.
  264. N. I. Anaraki, A. Sadeghpour, K. Iranshahi, C. Toncelli, U. Cendrowska, F. Stellacci, A. Dommann, P. Wick and A. Neels, Nano Res., 2020, 13, 2847–2856 CrossRef CAS.
  265. D. García-Lojo, E. Modin, S. Gómez-Graña, M. Impéror-Clerc, A. Chuvilin, I. Pastoriza-Santos, J. Pérez-Juste, D. Constantin and C. Hamon, Adv. Funct. Mater., 2021, 31, 2101869 CrossRef.
  266. T. Ehm, J. Philipp, M. Barkey, M. Ober, A. T. Brinkop, D. Simml, M. von Westphalen, B. Nickel, R. Beck and J. O. Radler, J. Synchrotron Radiat., 2022, 29, 1014–1019 CrossRef CAS PubMed.
  267. D. Radajewski, P. Roblin, P. Bacchin, M. Meireles and Y. Hallez, Lab Chip, 2023, 23, 3280–3288 RSC.
  268. J. D. Guild, S. T. Knox, S. B. Burholt, E. M. Hilton, N. J. Terrill, S. L. M. Schroeder and N. J. Warren, Macromolecules, 2023, 56, 6426–6435 CrossRef CAS PubMed.
  269. T. D. Turner, C. O’Shaughnessy, X. He, M. A. Levenstein, L. Hunter, J. Wojciechowski, H. Bristowe, R. Stone, C. C. Wilson, A. Florence, K. Robertson, N. Kapur and F. C. Meldrum, J. Appl. Crystallogr., 2024, 57, 1299–1310 CrossRef CAS PubMed.
  270. E. G. Kozyr, P. N. Njoroge, S. V. Chapek, V. V. Shapovalov, A. A. Skorynina, A. Y. Pnevskaya, A. N. Bulgakov, A. V. Soldatov, F. Pellegrino, E. Groppo, S. Bordiga, L. Mino and A. L. Bugaev, Catalysts, 2023, 13, 414 CrossRef CAS.
  271. T. C. Miller, M. R. Joseph, G. J. Havrilla, C. Lewis and V. Majidi, Anal. Chem., 2003, 75, 2048–2053 CrossRef CAS PubMed.
  272. A. A. Maurice, J. Theisen, V. Rai, F. Olivier, A. El Maangar, J. Duhamet, T. Zemb and J.-C. P. Gabriel, Nano Sel., 2022, 3, 425–436 CrossRef CAS.
  273. F. L. Olivier, S. M. Chevrier, B. Keller and J.-C. P. Gabriel, Chem. Eng. J., 2023, 454, 140306 CrossRef CAS.
  274. S. Youssef, D. Bauer, S. Békri, E. Rosenberg and O. Vizika, presented in part at the International Symposium of the Society of Core Analysts, SCA2009-17, Noordwijk aan Zee, The Netherlands, 27-30 September, 2009 Search PubMed.
  275. M. Andrew, B. Bijeljic and M. J. Blunt, Geophys. Res. Lett., 2013, 40, 3915–3918 CrossRef CAS.
  276. M. Andrew, B. Bijeljic and M. J. Blunt, Int. J. Greenhouse Gas Control, 2014, 22, 1–14 CrossRef CAS.
  277. M. Andrew, B. Bijeljic and M. J. Blunt, Adv. Water Resour., 2014, 68, 24–31 CrossRef CAS.
  278. T. Bultreys, M. A. Boone, M. N. Boone, T. De Schryver, B. Masschaele, L. Van Hoorebeke and V. Cnudde, Adv. Water Resour., 2016, 95, 341–351 CrossRef.
  279. H. P. Menke, B. Bijeljic, M. G. Andrew and M. J. Blunt, Environ. Sci. Technol., 2015, 49, 4407–4414 CrossRef CAS.
  280. P. Gajjar, J. S. Jørgensen, J. R. A. Godinho, C. G. Johnson, A. Ramsey and P. J. Withers, Sci. Rep., 2018, 89, 093702 Search PubMed.
  281. J. R. A. Godinho and P. J. Withers, Geochim. Cosmochim. Acta, 2018, 222, 156–170 CrossRef CAS.
  282. K. Singh, A. T. M. S. Huqe Muzemder, D. Edey, M. Colbert, J. Maisano and B. Shafei, Appl. Geochem., 2024, 167, 105980 CrossRef CAS.
  283. S. A. Mäkiharju, J. Dewanckele, M. Boone, C. Wagner and A. Griesser, Exp. Fluids, 2021, 63, 16 CrossRef.
  284. T. Bultreys, S. Van Offenwert, W. Goethals, M. N. Boone, J. Aelterman and V. Cnudde, Phys. Fluids, 2022, 34, 042008 CrossRef CAS.
  285. K. Tsuji, T. Emoto, Y. Nishida, E. Tamaki, Y. Kikutani, A. Hibara and T. Kitamori, Anal. Sci., 2005, 21, 799–803 CrossRef CAS PubMed.
  286. K. Nakano and K. Tsuji, J. Anal. At. Spectrom., 2010, 25, 562–569 RSC.
  287. K. G. McIntosh, J. A. Neal, P. Nath and G. J. Havrilla, X-Ray Spectrom., 2014, 43, 332–337 CrossRef CAS.
  288. R. B. Hammond, X. Lai, K. J. Roberts, A. Thomas and G. White, Cryst. Growth Des., 2004, 4, 943–948 CrossRef CAS.
  289. S. Dharmayat, R. B. Hammond, X. Lai, C. Ma, E. Purba, K. J. Roberts, Z.-P. Chen, E. Martin, J. Morris and R. Bytheway, Cryst. Growth Des., 2008, 8, 2205–2216 CrossRef CAS.
  290. S. Fouilloux, A. Désert, O. Taché, O. Spalla, J. Daillant and A. Thill, J. Colloid Interface Sci., 2010, 346, 79–86 CrossRef CAS PubMed.
  291. J. Huang, F. Deng, B. Günther, K. Achterhold, Y. Liu, A. Jentys, J. A. Lercher, M. Dierolf and F. Pfeiffer, J. Anal. At. Spectrom., 2021, 36, 2649–2659 RSC.
  292. P. Zimmermann, S. Peredkov, P. M. Abdala, S. DeBeer, M. Tromp, C. Müller and J. A. van Bokhoven, Coord. Chem. Rev., 2020, 423, 213466 CrossRef CAS.
  293. N. S. Genz, A.-J. Kallio, F. Meirer, S. Huotari and B. M. Weckhuysen, Chem.: Methods, 2024, 4, e202300027 CAS.
  294. V. C. Tidwell, L. C. Meigs, T. Christian-Frear and C. M. Boney, J. Contam. Hydrol., 2000, 42, 285–302 CrossRef CAS.
  295. A. Polak, A. S. Grader, R. Wallach and R. Nativ, Water Resour. Res., 2003, 39, 1106 CrossRef.
  296. T. J. Heindel, J. N. Gray and T. C. Jensen, Flow Meas. Instrum., 2008, 19, 67–78 CrossRef CAS.
  297. A. Sheppard, S. Latham, J. Middleton, A. Kingston, G. Myers, T. Varslot, A. Fogden, T. Sawkins, R. Cruikshank, M. Saadatfar, N. Francois, C. Arns and T. Senden, Nucl. Instrum. Methods Phys. Res., Sect. B, 2014, 324, 49–56 CrossRef CAS.
  298. M. Dierick, D. Van Loo, B. Masschaele, J. Van den Bulcke, J. Van Acker, V. Cnudde and L. Van Hoorebeke, Nucl. Instrum. Methods Phys. Res., Sect. B, 2014, 324, 35–40 CrossRef CAS.
  299. G. R. Myers, A. M. Kingston, T. K. Varslot, M. L. Turner and A. P. Sheppard, Appl. Opt., 2011, 50, 3685–3690 CrossRef PubMed.
  300. H. H. Shi, Y. Xiao, S. Ferguson, X. Huang, N. Wang and H. X. Hao, Lab Chip, 2017, 17, 2167–2185 RSC.
  301. M. L. Kovarik, P. C. Gach, D. M. Ornoff, Y. L. Wang, J. Balowski, L. Farrag and N. L. Allbritton, Anal. Chem., 2012, 84, 516–540 CrossRef CAS PubMed.
  302. R. Courson, S. Cargou, V. Conedera, M. Fouet, M. C. Blatche, C. L. Serpentini and A. M. Gue, RSC Adv., 2014, 4, 54847–54853 RSC.
  303. S. Guha, S. L. Perry, A. S. Pawate and P. J. A. Kenis, Sens. Actuators, B, 2012, 174, 1–9 CrossRef CAS PubMed.
  304. Y. Qin, J. E. Kreutz, T. Schneider, G. S. Yen, E. S. Shah, L. Wu and D. T. Chiu, Lab Chip, 2022, 22, 4729–4734 RSC.
  305. K. Sugioka, J. Xu, D. Wu, Y. Hanada, Z. Wang, Y. Cheng and K. Midorikawa, Lab Chip, 2014, 14, 3447–3458 RSC.
  306. A. Liga, J. A. S. Morton and M. Kersaudy-Kerhoas, Microfluid. Nanofluid., 2016, 20, 1–12 CrossRef CAS.
  307. R. Su, F. Wang and M. C. McAlpine, Lab Chip, 2023, 23, 1279–1299 RSC.
  308. N. J. Szymanski, B. Rendy, Y. Fei, R. E. Kumar, T. He, D. Milsted, M. J. McDermott, M. Gallant, E. D. Cubuk, A. Merchant, H. Kim, A. Jain, C. J. Bartel, K. Persson, Y. Zeng and G. Ceder, Nature, 2023, 624, 86–91 CrossRef CAS PubMed.
  309. C. W. Coley, D. A. Thomas, J. A. M. Lummiss, J. N. Jaworski, C. P. Breen, V. Schultz, T. Hart, J. S. Fishman, L. Rogers, H. Gao, R. W. Hicklin, P. P. Plehiers, J. Byington, J. S. Piotti, W. H. Green, A. J. Hart, T. F. Jamison and K. F. Jensen, Science, 2019, 365, eaax1566 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.