Skip to main content
NIST Author Manuscripts logoLink to NIST Author Manuscripts
. Author manuscript; available in PMC: 2025 Jan 22.
Published in final edited form as: Nat Rev Chem. 2019 Apr 26;3(5):305–314. doi: 10.1038/s41570-019-0097-z

Chemistry from 3D printed objects

Matthew R Hartings 1, Zeeshan Ahmed 2
PMCID: PMC11753264  NIHMSID: NIHMS1533764  PMID: 39845774

Abstract

3D printing technology has started to take hold as an enabling tool for scientific advancement. Born from the marriage of computer-aided design and additive manufacturing, 3D printing was originally intended to generate prototypes for inspection before their full industrial production. As this field has matured, its reach into other applications has expanded, accelerated by its ability to generate 3D objects with complex geometries. Chemists and chemical engineers have begun to take advantage of these capabilities in their own research. Certainly, the most prominent examples of this adoption have been the design and use of 3D printed reaction containers and flow devices. The focus of this Review, however, is on 3D printed objects, the chemical reactivities of which are of primary interest. These types of objects have been designed and used in catalytic, mechanical, electronic, analytical and biological applications. Underlying this research are the efforts to add chemical functionality to standard printing materials, which are often inert. This Review details the different ways in which chemical reactivity is endowed on printed objects, the types of chemical functionality that have been explored in the various printing materials and the reactions that are facilitated by the final printed object. Finally, the Review discusses new avenues for the development and further sophistication of generating chemically active, 3D printed objects.


The first 3D printing process was developed by Hideo Kodama in the early 1980s1 (FIG. 1a). He was able to generate a complete 3D object from 2D cross sections, successively formed on top of one another, using photo-hardening polymers and UV light. Kodama’s apparatus consisted of a print bed, which contained a liquid polymer resin (an unsaturated polyester), a crosslinking agent, an initiator and a photosensitizer. A solid structure was produced by either raster-scanning an optical fibre over the print bed or placing a mask over the bed and fully illuminating the exposed surface. The growing object was then lowered into the bed, exposing unset liquid resin, and a new solid layer was formed on top until the object was fully fabricated.

Fig. 1 |. Examples of 3D printed objects.

Fig. 1 |

a | One of the original pieces printed by Kodama in the early 1980s1. b | Typical laboratory supplies generated with a 3D printer11. c | A 3D printed fluidic device23. d | 3D printed ‘reactionware’ for performing a multistep, organic synthesis and purification28. Part a is reproduced with permission from REF.1, AIP Publishing. Part b is reproduced from REF.11, Springer Nature Limited. Part c is reproduced with permission from REF.23, Wiley-VCH. Part d is adapted with permission from REF.28, AAAS.

Current 3D printing technologies have advanced, in great part because of computer generation of 3D structures and control over the printing processes29. Many modern printing technologies, though, still use the chemical foundation exploited by Kodama to build his device. 3D printing techniques can be sorted into two distinct classifications. The first class includes methods, like Kodama’s, in which an object is generated from a reservoir of patternable or polymerizable materials. The second class includes processes in which a material is extruded through a nozzle onto a platform. Horizontal movement of the nozzle and vertical movement of the platform during extrusion defines the final shape of the 3D printed object.

Although 3D printing was initially used by designers to evaluate industrial prototypes, the widespread use and popularity of 3D printing have grown because of the enthusiasm of at-home users. These makers most often use extrusion-based printers, which tend to be the least expensive to own and operate. An entire online community has evolved in which makers share their object designs and provide tips for managing tricky or complicated prints. Not to be left out of the fun, scientists are also finding ways to use 3D printing for their work. Researchers can print supplies for their laboratories10,11 (FIG. 1b). In some cases, objects cost less to print than they would to purchase from laboratory supply companies. An added benefit to this approach is that there is no wait time for item orders to be processed and shipped. Scientists have also designed objects that fill a very specific niche for which there are no commercially available products12. Finally, there are many examples of 3D printed replicas that scientists use to visualize their research. 3D printed organisms, organs and chemical structures (both large and small) can be used to educate and engage non-specialists as well as facilitating deeper understanding and insights from the scientists who study them.

Chemists and chemical engineers have begun to move past printing replicas to printing objects that facilitate research in new and interesting ways. Custom reaction vessels have been generated in an attempt to study and optimize reaction conditions and product yields68,12,13. One example of this type of work includes the preparation of flow reactors1423 (FIG. 1c). These types of devices can exhibit flow circuitry designed to accommodate specific reaction kinetics (precise reaction times) and thermodynamics (safely facilitating highly exothermic reactions). Flow reactors can also accommodate multicomponent reactions in which each component is added at a specific time point during the reaction. Perhaps most relevant to the capabilities of 3D printing are reactors that employ mixing chambers with unique geometries, which might otherwise be difficult or impossible to generate2427. Along with flow devices, larger reaction vessels have also been printed12,2831. In one example of this type of ‘reactionware’, Cronin and colleagues generated a connected network of containers that facilitated the synthesis of baclofen (a muscle relaxant)28 (FIG. 1d). Their printed container facilitated three reactions along with several extractions, filtrations and isolation steps required to generate intermediates and final product.

In the above examples, the reaction containers necessitate the printing material to be chemically resistant. Although these printed devices facilitate chemistry, the materials from which they are made are not actively involved in the chemistry of the reactions. There has been a growth in the number of studies in which scientists produce printed objects that are chemically active. In this Review, we focus on ways in which chemical functionality can be added to 3D printing materials and highlight areas where we feel these materials can make an impact on future research efforts. Specific equipment and materials are identified in this Review in order to describe the experimental procedure adequately. Such identification is not intended to imply endorsement by the National Institute of Standards and Technology, nor is it intended to imply that the materials or equipment identified are necessarily the best available.

Printing processes

There are a number of relevant publications that fully detail the different styles of 3D printing, instrumentation, their material requirements, post-printing processing steps and capabilities28. Here, we describe some of these aspects mainly to highlight how the maintenance or addition of chemical reactivity can be facilitated (FIG. 2).

Fig. 2 |. Additives that imbue chemical reactivity in 3D printed objects.

Fig. 2 |

a | Polymers in 3D printing materials can have reactive side chain or backbone functional groups. b | Different types of particle can be added to 3D printing materials to facilitate chemical reactivity in printed objects. These particles include platelets, nanoparticles, microparticles, fibres, 2D materials and porous particles such as metal organic framework and zeolitic materials.

FIGURE 2 shows the types of materials from which a 3D printed object can derive reactivity. Polymers are the basis for nearly all 3D printing processes. Relevant functionalities are found in both the polymer backbone and side chains32,33, which can be used as they are or altered after printing. Nanoparticles and microparticles have also imbued printed objects with reactivity31,34,35. Particles can be used as the primary component of a 3D printing paste35, included as a secondary component to the printing material34 or incorporated onto the surface of an already printed object31. Platelet-shaped particles and high-aspect ratio particles have been primarily used to alter the mechanical properties of the printing material36 but can also facilitate interesting chemistry, which is the case for montmorillonite K10 and its ability to catalyse a number of organic transformations31. 2D materials (such as graphene and graphene oxide), normally incorporated before printing, have been used as rheological modifiers to enable printing and for their specific reactivities37,38. Metal organic frameworks (MOFs) and zeolite particles, valued for their high internal surface areas, are incorporated before printing or synthesized on the surface of a printed object32,3941.

In this Review, we explore the different printing techniques and commonly used materials that enable novel chemistry from 3D printed objects (FIGs 3,4). We detail the ways in which these materials can be modified and the interesting chemical processes they facilitate. The cost of purchasing a 3D printer or the difficulty in producing modified printing materials may be an entry barrier to this field. However, many companies now sell customer-defined objects printed from a variety of materials, providing a solution for those that are curious but do not want to incur some of the cost or effort required.

Fig. 3 |. Overview of reactive materials that can be used in various 3D printing techniques.

Fig. 3 |

a | Filaments used in fused deposition modelling (FDM) are made from any polymer composite that can be extruded at high temperatures and quickly solidified upon cooling. Chemical reactivity in FDM materials can come from the primary polymer or from additives (blended polymer, blended nanoparticle or reactive moiety added after the printing process). Care must be used in modifying the filaments before printing, as the high temperatures required for extrusion can degrade some additives. b | Robocasting, direct ink writing (DIW) and inkjet 3D printing all use liquid-based inks. Chemically active species can be part of the printed ink or can be added after printing (in the case of robocasting and DIW). c | Stereolithography (SLA), digital light processing (DLP) and two-photon 3D printing generate objects through photopolymerization of reactive resins. These resins can incorporate reactive functional groups (orthogonal to the photochemistry) or reactive particles or can be functionalized after printing. d | 3D powder printing involves using a printed binder liquid to glue together particles in a powder. The chemical reactivity of these objects can come from the powder or from the deposited ink. e | For objects made with metal or ceramic powders, a final sintering step is required that will degrade any organic functionality from the binder liquid. In sintering and melting-based 3D printers, a polymer or metal powder is locally heated to join individual particles into a larger object. The powder can be the basis for the desired chemical reactivity, or the object can be functionalized after the printing process. Metal sintering printers have a higher cost than those that use polymers.

Fig. 4 |. Technical details of different 3D printing approaches.

Fig. 4 |

3D printing methods to imprint chemical activity to an object are compared. DLP, digital light processing; FDM, fused deposition modelling; SLA, stereolithography.

Fused deposition modelling

In fused deposition modelling (FDM), a thermoplastic filament is fed through a nozzle that is heated above the glass transition temperature of the polymer42,43. The polymer flows through the nozzle and solidifies after being deposited on the print bed. FDM printers are the most commonly used by hobbyists. A number of commercially available polymer filaments in a variety of colours with a variety of mechanical and electronic properties can be used for FDM. The most typically used filaments are composed of acrylonitrile butadiene styrene (ABS) and polylactic acid (PLA). Other FDM plastics include high-impact polystyrene (HIPS), polyvinyl alcohol (PVA), polypropylene (PP), polyethylene terephthalate (PET), polyethylene terephthalate glycol (PETG), polycarbonate (PC), polycaprolactone (PCL), polyethylene co-trimethylene terephthalate (PETT) and nylon. There are very few easily accessible functional groups on these polymers; the nitrile groups on the ABS side chains, the hydroxyl groups on PVA and the ester backbone of PLA are the rare examples (FIG. 3a).

ABS is a staple in injection-moulded consumer products. As such, there are a number of commercially available coating technologies that treat ABS for use in different applications44,45. These processes have yet to be explored as a way to tailor the reactivity of 3D printed ABS objects. Additionally, the organic reactions for transforming nitriles are also well known but have not been adapted for modifying 3D printed objects46. Both of these approaches offer pathways for imbuing ABS objects with chemical reactivity.

PLA has also been used to generate objects for research purposes. For example, the incubation of PLA-based biological scaffolds in aqueous systems results in the hydrolysis of the ester bond of the polymer47. As discussed below, this hydrolytic reactivity plays an important role in the success of the scaffold, as it supports cell growth and tissue development. As with modification of ABS objects, the transformation of esters has yet to be explored as a means to add specific reactive functionalities to PLA.

Several researchers have manufactured their own filaments to print objects for very specific applications. One approach in this direction is to synthesize custom polymers. Boydston and co-workers produced PCL filaments by reacting caprolactone monomers in the presence of caprolactone-modified spiropyran48. The resulting polymer was formed into a printable filament using a common extruder. Some researchers have incorporated functionality into their filaments by mixing traditional 3D printing polymers with other materials. Several groups have solvent cast suspensions of inorganic nanoparticles and polymers and generated printable filaments from the evaporated films34,40,41,49. Despite the ease of this approach, there is a risk of phase separation between the printing polymer and the additive during evaporation. Another approach involves the use of the blending instrumentation regularly employed in the polymer industry. Twin-screw extruders are used to compound dyes and polymers to generate commercially available coloured filaments. Similarly, these instruments can blend a printing polymer with other types of additives. We used this approach to generate ABS–MOF composites39. During this process, we flooded the system with nitrogen to ensure the stability of the organic linkers in the MOF at the high temperatures required for blending. MOF particles have also been grown on printed thermoplastic objects and tested for their ability to remove dye molecules from water32,50. Liu and co-workers have synthesized HKUST-1 (a MOF composed of copper ions and 1,3,5-benzenetricarboxylic acid) on an ABS object32, whereas Zhang and co-workers synthesized the same MOF on a PLA object.50

One issue with the use of FDM-based approaches for 3D printing chemically active objects is that the polymers tend to have low porosity. This leaves the surfaces exposed by the object as the best sites for modification if the goal is to produce efficient catalysts. One approach for overcoming this limitation is to generate pores within a 3D printed object. Porogens are materials that have different solubility from the primary structural polymer51. For composites containing a printable polymer and reactive insert, the choice of porogen must be made such that the initial blending and final solvent treatment do not affect the distribution of the reactive insert in the structural polymer. This approach has been successfully employed to produce a commercially available filament that is composed of a rubber-elastomeric polymer and PLA. Pores are generated by soaking printed objects in water, solubilizing and removing the PLA.

Another issue that arises in FDM printing is that not all thermoplastics print as easily as others. As an example, PLA does not require as many controls on external parameters during printing as ABS. PLA can be printed directly from the heated nozzle onto an unheated platform. ABS, however, contracts as it cools. During printing, this can lead to object warping. We have found that printing on a heated platform in an insulated chamber works best for ABS34,39,41. Issues with printing quality (such as, changes in thermo-physical properties, rheology, warping and others) complicate laboratory-manufactured filaments of pure polymers, polymer blends and polymer composite materials.

Robocasting and direct ink writing

In robocasting or direct ink writing (DIW), an ink (often a gel or paste) is extruded through a syringe onto a platform in a similar manner to the FDM process4. The ink is composed of a solvent and either a gelling polymer or paste-producing nanoparticles. Often, these inks are made by mixing multiple components to generate optimal rheological properties52. These properties include the ability to be extruded through a syringe needle or tip under an applied pressure and to maintain a given shape after extrusion when no force is being exerted. There are several procedures that can be used to achieve these results. Similar to FDM, a solid gel can be heated until it exhibits flow and extruded onto a cooled platform. Gelatin-based suspensions can be printed in this manner53. Another approach is to create a solid-like suspension that flows only under periods of sheer stress. Inks can be based on hydrogels, in which the polymer forms a web-like solid network within the solvent54. Inks can also be based on inorganic nanoparticles, in which particle–particle interactions (and not particle–solvent or solvent–solvent interactions) dictate rheological properties37. To achieve the high particle concentrations for these inks, it is often necessary to use orbital mixers for blending.

The great benefit of DIW is that inks can be generated on demand. Users have free latitude to include reactive molecules or particles as they see fit. These have included polymers5557, proteins58, graphene38, graphene oxide37, nanoparticles and microparticles (of clays and ceramics59,60, metal oxides35, quantum dots61, metals62,63, zeolites64 and MOFs65) (FIG. 3b). The use of metal oxide particles has been especially notable in the production of catalytic 3D printed objects.

The challenge to DIW is that printed objects need to have stability within the environments in which they are to be used. The structures generated with DIW require extra processing steps to achieve the mechanical properties of objects produced by FDM. In most cases, the solvent needs to be carefully removed from the object, usually using gentle heating and lyophilization60. Methods to optimize the mechanical properties of these objects vary depending upon the materials used. Some combination of inorganic particles and polymers will generate stability66. Some polymer-containing systems can be crosslinked67,68. Metal oxide-based systems can be sintered, although any organic molecules present will be carbonized60. In some instances, carbonization can be a feature instead of an issue. Travitzky and co-workers and Hutmacher and co-workers have generated graphitic structures using DIW with inks that contain starch and by carbonizing the resulting object69,70.

Once a stable object has been generated, there is an array of options for chemical functionalization. Objects can be coated with thin films71, and molecular components, ranging from molecular transition-metal catalysts72 to biologically relevant small molecules and proteins73,74, can be covalently attached.

Inkjet printing

Inkjet printers can manufacture small-scale (micrometres to millimetres) 3D printed objects7578. In this case, inks are composed of a solvent along with a structure-forming material (FIG. 3). Liu and co-workers, for example, generate their inks from ethanol, acetic acid, hydrochloric acid and structure-forming materials (a triblock co-polymer and metal ions)77. Another example of an ink used in this kind of printing from Parkinson and co-workers includes a metal salt in a water–glycol mixture78. A general schematic of the inks is shown in FIG. 3b. As opposed to the inks used in DIW, those required for inkjet printing do not need to exhibit thixotropic properties, typical of thick fluids that change their viscosity under an applied force. In fact, these inks are optimized with lower viscosities and fast evaporating solvents. A benefit of using inkjet printing is that it is easy to use multiple inks during the same printing process. Most of the chemically reactive objects produced by inkjet printing so far are new catalytic materials6,7678. The ability to use multiple ink wells allows for mixing metals and metal oxides at various ratios for testing reactivity. There are other printing technologies that combine inkjet deposition with controlled material curing steps. One example of this is the PolyJet technology from Stratasys, which combines inkjet printing and photopolymerization processes (described in the next section) into a single printer. This technology can generate large (with sizes on the order of hundreds of centimetres) and intricate structures (with features as small as 100 micrometres) from a variety of polymers and polymer composites from inkjet printing processes.

SLA, DLP and two-photon 3D printing

Stereolithography (SLA)7981, digital light processing (DLP)82 and two-photon83 3D printing use photochemistry to generate solid polymers from a liquid resin (FIG. 3c). SLA and two-photon printers use lasers, whereas DLP uses projected light, structured into a 2D image. DLP tends to generate objects more quickly than SLA, which requires a laser to be raster scanned across the surface of the resin. Continuous light interface production (CLIP), an advanced form of DLP, uses oxygen diffusion to control the polymerization process and speed printing84. Two-photon 3D printing polymerizes resin only at the focus of a laser passed through a microscope objective. This process can generate objects with the smallest print resolution (on the order of 100 nm)83 but is limited by the object size and time required for printing.

Resins used in SLA, DLP and two-photon 3D printing are composed of acrylic, epoxy or urethane-based molecules and can be modified both before printing and after printing to achieve the desired chemical functionalities8588. Resins with designed chemical reactivity orthogonal to the polymerization process can be synthesized and used in the same manner as the commercially available resins. Polyethylene glycol (PEG) dimethyl acrylate resins have been used to generate biocompatible scaffolds89,90. Boydston and colleagues have developed special resins that can generate flexible printed structures88. There is no reason why similar modifications that incorporate catalytic functionalities could not be used for the production of chemically reactive objects.

Polymer composites can also be generated during photopolymerization. Carbon, the company that licenses the CLIP technology, generates composites that contain thermoset polymers within the polymerized scaffold. Post-printing heat treatments ultimately dictate the mechanical properties of the printed object. Gao and colleagues have incorporated nanoscale hydroxyapatite into their printed objects to facilitate the regeneration of bone tissue91. Hensliegh and co-workers have used SLA processes to produce polymer–graphene oxide composite structures81. Several groups have used resin curing to produce printed polymer–MOF composites92,93.

3D powder

For 3D powder printing, an inkjet is raster scanned across a bed of powder94,95. The ink solution contains a chemical glue that binds the powder particles together (FIG. 3d). As the object is generated, the platform is lowered, a new layer of powder is added on top, and the next layer of the object is formed through the application of binder ink. A number of polymer powders can be used in this process as can ceramics and metals. The metal and ceramic objects built this way require sintering to achieve full stability. Polymeric objects from these types of printer have been used in a number of medically related applications70,9698.

Sintering and melting

Laser sintering and laser melting techniques use high-power laser sources, scanned over a bed of powder, to selectively sinter the powders into a solid object or melt the powders, which form a solid object on cooling (FIG. 3e). New layers are added in a similar manner to the 3D powder printing technique. There are laser sintering and melting instruments that are designed for either polymers (polyamide 12 is the most widely used) or metal or ceramic particles25,94.

Chemical reactivity

Chemical functionality can be manifested in a number of ways, and catalytic properties come immediately to mind. In the next sections, we discus 3D printed catalytic objects but also touch upon other, perhaps subtler, types of chemical reactivity. Molecular recognition and binding, signal transduction and bond lability are all types of chemical functionality that are found and applied in 3D printed objects. We detail examples of these types of chemical properties for applications in catalysis, molecular storage, electronics, signal transduction and tissue culture.

Chemical catalysis

Catalytic objects are ideal targets for researchers using 3D printing79 (FIG. 5a). The ability to define a printed material gives the researcher control over the reactions that they wish to catalyse. Furthermore, the ability to define object geometry can lead to structures that optimize the flow of reactants through a printed reactor. Because of these capabilities, 3D printing has the potential to increase the efficiencies and decrease the cost for many industrial processes.

Fig. 5 |. Examples of printed object geometries that exploit chemical functionality.

Fig. 5 |

A | Catalyst architectures (left: woodpile; right: porous helix) increase the exposed surface area of the printed object and induce turbulent flows in reactive devices. B | 3D printed objects made of a stimulus-responsive wax-based polymer. These polymers swell according to the duration of their light exposure and were used to print different planar layouts (inserts). Exposing the different parts of the printed layouts (light areas correspond to short exposures, and dark areas correspond to long exposures) leads to complex 3D shapes: a helix (top left) and a tulip flower (bottom left). The objects can be deformed at high temperature (top and bottom centre), but the initial shape can be recovered by cooling (top and bottom right)107. C | 3D printed metal-organic framework material (MOF) monoliths for gas storage. Parts Ca and Cb show scanning electron microscopy (SEM) images of MOF-74(Ni) monoliths, and parts Cc and Cd show SEM images of UTSA-16(Co) monoliths65. D | 3D printed materials for sensing mechanical distortions, before (top) and after (bottom) elongation48. E | Design targets for tissue engineering featuring gradients of mechanical strength and concentrations of growth factors are shown. Part B is adapted with permission from REF.107, Wiley-VCH. Part C is adapted with permission from REF.65, ACS. Part D is adapted with permission from REF.48, ACS.

One of our goals, when first thinking about generating chemically active 3D printed objects, was to create a printing material that could potentially have a low entry barrier for casual users. ABS and PLA-based FDM filaments seemed to be ideal for this purpose. We solvent cast TiO2 nanoparticles and ABS from an acetone suspension34. The resulting TiO2–ABS composites were formed into filaments, printed and evaluated for photocatalytic activity. The printed objects were capable of degrading a dye when they were in direct sunlight, but they showed no activity when stored in the dark.

Although the approach we proposed offers broad accessibility, the composite polymer–catalyst material is not optimized to efficiently facilitate catalysis. For a reactive object, the exposed surface plays the largest role in supporting catalysis. In our approach, the reactive nanoparticles within the polymer composite are not optimally exposed to the analytes of interest. To this end, some researchers have printed reaction containers using FDM and modified the surfaces of these containers with reactive moieties. Cronin and colleagues printed a set of connected reaction containers with polypropylene12,31. In one chamber, the surface was modified with montmorillonite K10, applied as a paste of acetoxysilicone. In another chamber, the surface was modified with Pd on carbon catalysts, applied in a similar manner. The reactionware facilitated a Diels–Alder reaction (catalysed by the montmorillonite K10) and a reduction (catalysed by the Pd on carbon). In another experiment, the same group modified an object, printed with polypropylene, with silver paint followed by a sputtering of gold. They used these reaction containers to support the electrolysis of water99.

Perhaps the most used technique for generating catalytic 3D objects involves the direct print of the catalytic material through DIW or inkjet printing. In both cases, inks are composed of reactive particles suspended in solvent. After printing, the solvent is removed and the object sintered. For the cases in which inkjet printing is used, object formation usually occurs using some sort of non-reactive glass as a substrate. DIW can generate freestanding metal and metal oxide objects that are larger in size than those made using inkjets. Catalytic monoliths have been produced and discussed in detail elsewhere79, but we report a few examples below.

Michorczyk and colleagues used a DLP printer to produce a polymer-based template of a catalytic monolith100. They filled the template with a manganese-doped sodium tungstate paste. Burning off the polymer and sintering the paste, they were left with a monolith structure that contained open channels in which to flow reactants. They tested these catalytic monoliths for the conversion of methane into larger hydrocarbons at 800 °C. Sotelo, Gil and colleagues used DIW to print a catalytic monolith from a copper-doped Al2O3 paste62. The sintered object was then used to perform a number of copper-catalysed organic transformations. Zhu and co-workers generated a gold monolith with nanoporous features by first printing an ink composed of a polymer and a gold–silver alloy63. Sintering and dealloying processes generated a final gold object, which allowed for fast mass transport and catalysis in the oxidation of methanol to methyl formate. Finally, Gil, Coelho and colleagues used DIW to print silica pastes. After sintering, they used silane chemistry to modify the surface of the monoliths with either a Cu(i)-based or a Pd(ii)-based molecular catalyst72. Placing both monoliths in a reaction container, they were able to perform a multicomponent reaction that included a copper-catalysed alkyne–azide cycloaddition and a palladium-catalysed cross coupling.

Catalytic architectures can also be generated using photocuring resins. Yang and co-workers used a bromine-containing photocuring initiator85. After the printing process, they used the initiator incorporated within the object as a site for facilitating atomic-transfer radical polymerization. Slowing and colleagues printed objects with resins containing a number of functional groups (carboxylic acids, amines and copper carboxylate) along with the vinyl groups necessary for curing87. These secondary functional groups were active for Mannich, aldol and cycloaddition reactions.

Stimulus sensing and response

There are many examples of objects printed from stimulus-responsive materials, activated through chemical, mechanical and electrical changes, to name a few18,48,56,75,101103. The ability to generate programmable chemical composites coupled to the unique geometry-creating capacity of 3D printing can lead to a variety of responsive objects. Highlighting the marriage of geometry and material is the printing of photonic devices and optical waveguides49,75,104.

Of all stimulus-responsive objects, electronically active objects have perhaps garnered the most interest56,82,105. This is evident from the variety of commercially available, conductive FDM filaments. Several forms of conductive materials have been explored by researchers. A silver nanoparticle–silicone composite was printed with robocasting and used as the sound-detection and signal transduction medium of a printed bionic ear101. Conductive nanoparticles have been used in FDM, DIW, inkjet, sintering and melting-based printers. The research groups lead by Travitsky and Hutmacher have used robocasting to print polymeric materials (including starch), which are then carbonized. Graphene and graphene oxide-based materials have garnered much interest as well69,70. Graphene oxide has been incorporated into ABS for FDM printing using solvent exchange106. Graphene oxide has been used as the primary rheological modifier for generating inks to be used in DIW37. Graphene-based aerogels have been generated by DIW of graphene inks within a non-solvent matrix38.

4D printing is a name that has been given to the design and printing of objects that can change shape in response to an external stimulus107,108 (FIG. 5b). Shape control in these objects occurs in two ways: mechanically deformed objects revert to their original shape upon changes in temperature, and hydrogels swell in response to chemical stimuli. Dunn, Qi and co-workers have printed objects composed of both hydrogels and temperature-controlled, shape-memory polymers109. Zhao and collaborators have printed magnetic domains within soft materials using DIW, aligning the magnetic particles within the printed material with an applied field107. The result is a mechanically active material that responds to changes in the local magnetic field.

Molecular storage and separation

Substantial advances have been made on materials that can store gases at higher densities than possible in traditional compressed gas cylinders. MOF and zeolitic materials are in the midst of a renaissance; options for combinations of metal ions and organic linkers, theoretical modelling of active sites and crystalline defects and the allure of applications in storage, sensing and catalysis have caught the imagination of materials scientists110. One of the issues for full deployment of MOFs is that they are synthesized as powders. For applications based solely on storage, the powder-based nature of MOFs is not an issue. However, incorporation of MOFs (and their functionality) into other devices will require that they be processible. Blending MOFs in polymers can have the added benefit of preventing MOF degradation owing to humidity39.

There has been an effort to combine MOF particles with 3D printing technologies. Liu and collaborators have grown HKUST-1 crystals on 3D printed ABS scaffolds32. Several researchers have generated FDM filaments by solvent casting to produce polymer–MOF composites3941. Erikson has been able to display printing with a 50% blend of MOF and polymer40. As some MOFs are environmentally sensitive, we have found that long exposure times during solvent casting can degrade MOF particles during composite preparation41. For industrial preparation of coloured filaments, polymers need to undergo a melt-blend process with dye powders in twin-screw extruders. We have found this process to be effective in generating polymer–MOF filaments while protecting the MOF from ambient humidity by operating under a constant flow of nitrogen39. Furthermore, the MOF composites, which retain their gas storage capacity, can be incubated in water without any loss of crystalline structure. Rezaei and colleagues have generated DIW inks with 80% MOF loading and printed MOF monoliths that show absorption capacity similar to the unprocessed powder65 (FIG. 5c). Additionally, Chin and co-workers have used SLA to magnetically align different MOF crystals within a printed structure92.

Electrochemical and information storage

Objects have been printed in order to explore new ways to generate materials and architectures for electrochemical storage111113 (FIG. 5d). A recent review has highlighted many of these efforts113. Worsley, Li and co-workers have used DIW of graphene–MnO2 composite materials to print pseudocapacitive electrodes112. Sans and colleagues have used SLA to print objects that contain ionic liquids and polyoxometallates111 and tested their photochromic properties to reversibly store information.

Objects for use in tissue engineering

In many ways, researchers printing objects for tissue engineering (TE) have been at the forefront of the effort to print chemically reactive materials2,90,114. For TE, materials must have appropriate structure (on the macroscale, microscale and nanoscale), provide the possibility of nutrient transport and, importantly, facilitate appropriate cell–matrix interactions115. The ability to remodel tissue matrix is a critical factor that determines the ultimate success of cell growth and tissue development116. Efficient remodelling requires both cell attachment to and cleavage of the matrix material. That is, the chemical functionalities that must be present in any 3D printed TE scaffold are binding moieties that attach to cell surface proteins and structural integrity that can be enzymatically degraded and reprogrammed. For this reason, the primary components of many printed TE scaffolds are proteins (collagen, elastin, fibrin, gelatin, silk fibroin and decellularized extracellular matrix)114. These scaffolds are generally produced using robocasting, although inkjet printing and other droplet-based printing methods are also used.

One of the great benefits of these printing styles is that the ink formulation is user defined and is often easy to manufacture. As such, multiple inks can be printed during a single print job, resulting in programmed molecular or cellular gradients along a printed matrix (FIG. 5e). Composite inks that incorporate a number of useful functional materials can be generated at the researcher’s discretion. Importantly, these methods are also gentle enough to allow printing of materials that contain living cells.

Microcarriers are micrometre-scale particles that support cell adhesion and can facilitate the expression of specific cellular phenotypes. Levato and colleagues printed gradients of microcarrier ink to support the growth and location-specific differentiation of mesenchymal stem cells in a manner that mimics the gradient of cartilage composition in subchondral bone117.

For some tissue supports or medical implants, a sturdier support structure is needed than offered by the hydrogels generated with robocasting or droplet-based printing techniques. 3D powder-based printing allows for binding inks that contain biologically active materials. Metal or polymer particle sintering or melting and FDM methods can also generate the types of scaffold needed, but these objects must be supplemented with some sort of coating that can facilitate cell adhesion and differentiation91.

Specific growth factors and nutrients are also important constituents for an effective TE scaffold. Bone growth, supported by 3D printed objects, progresses much better in the presence of hydroxyapatite118,119. To that end, Zhang and colleagues have generated objects using SLA and a resin that contains dimethacrylated PEG and nanoscale hydroxyapatite particles118,119. The SLA process has a higher resolution than robocasting, which is ideal for the generation of narrow channels for nutrient flow.

Outlook

Research on chemically active 3D printed objects is in its early stages. The degrees of freedom in these objects include reactive moieties, matrix materials and object geometry. There are countless possibilities for generating reactive structures from 3D printing. Many advances will come through optimization of these three variables for producing objects that satisfy the particular needs of an application.

One of the aims of 3D printing technology development is to create controlled geometries over multiple length scales (nanoscale, microscale and macroscale). Resolutions vary by the type of printing process. Two-photon polymerization enables the smallest print feature sizes, on the order of hundreds of nanometres, but cannot produce large objects in reasonable print times. Other printing techniques enable the generation of features of more typical sizes, on the order of hundreds of micrometres. In TE applications, programmed control over chemical placement and structural geometry over these length scales are crucial for the success of printed scaffolds.

3D printed objects tend to have homogeneous composition. The generation of chemical composition or mechanical strength gradients along the length of an object is another area of current interest. Gradients of catalysts along a flow channel can facilitate entire reaction sequences instead of individual molecular transformations. Gradients of nutrient availability, molecular recognition sites and matrix mechanical strength are important factors for cellular differentiation and tissue development.

To date, most 3D printing is performed to generate well-defined objects with clean and smooth surfaces. To facilitate maximal chemical functionality, however, materials must display controlled porosity. There have been several studies that focused on generating programmed porosity in reactive objects63,120. It seems safe to say, however, that much work remains to be done in modelling and designing geometries that best serve a needed application.

Finally, we see great potential in combining multiple application modalities within an individual 3D printed object. Catalytic devices can be designed that facilitate reaction cascades and isolation of reaction products. A heterogeneous photocatalyst, composed of multiple metal oxide materials, can be generated to support reaction chambers in which the wavelength of light dictates which catalyst is activated. Waveguides can be printed with spatially resolved sensing modalities that can communicate both the presence and the location of analytes. Tissue scaffolds that support differentiation while also reporting development parameters can provide real-time analysis of implanted devices.

The future applications of chemically active, 3D printed objects abound. The democratization of manufacturing capabilities is bound to not only excite scientific curiosity across disciplines but also unleash users’ imaginations, opening new venues in the chemical sciences with applications far beyond.

Initiator.

A chemical needed to initiate polymerization. In some instances, initiators do this by generating radicals under mild conditions.

Photosensitizer.

Upon absorbing light, a photosensitizer can cause a change to a nearby chemical. In some instances, photosensitizers are used to generate radicals on initiators.

Printing paste.

A viscous fluid made from particles suspended in a solvent.

Thermoplastic filament.

A polymer that displays malleability at higher temperatures and is a solid at lower temperatures. For the purposes of fused deposition modelling 3D printing, a thermoplastic must be able to extrude through a nozzle at elevated temperatures.

Extruder.

An instrument that forces a fluid through a nozzle. For the purpose of 3D printing, the fluid can be a thermoplastic (fused deposition modelling), a gel or a paste (robocasting and/or direct ink writing).

Twin-screw extruders.

Devices used to homogeneously blend a polymer with another substance. During this blending process, the polymer is heated above its melting or glass transition temperature while two screws, which interpenetrate one another, continuously mix the components.

Hydrogels.

Water-based substances with increased viscosity caused by the interaction of macromolecules within the mixture.

Sintering.

The act of changing a powder into a solid material through application of heat and pressure without completely melting the powder.

Melt-blend.

The homogeneous mixture of a thermoplastic and other material (another thermoplastic, inorganic nanoparticle, polymer particle, and so on) made at elevated temperatures capable of melting the matrix thermoplastic.

Footnotes

Competing interest

The authors declare no competing interests.

References

  • 1.Kodama H Automatic method for fabricating a three-dimensional plastic model with photo-hardening polymer. Rev. Sci. Instrum 52, 1770–1773 (1981). [Google Scholar]
  • 2.Chia HN & Wu BM Recent advances in 3D printing of biomaterials. J. Biol. Eng 9, 4 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Peltola SM, Melchels FPW, Grijpma DW & Kellomaki M A review of rapid prototyping techniques for tissue engineering purposes. Ann. Med 40, 268–280 (2008). [DOI] [PubMed] [Google Scholar]
  • 4.Truby RL & Lewis JA Printing soft matter in three dimensions. Nature 540, 371–378 (2016). [DOI] [PubMed] [Google Scholar]
  • 5.Wang X, Jiang M, Zhou ZW, Gou JH & Hui D 3D printing of polymer matrix composites: a review and prospective. Compos. Part B 110, 442–458 (2017). [Google Scholar]
  • 6.Hurt C et al. Combining additive manufacturing and catalysis: a review. Catal. Sci. Technol 7, 3421–3439 (2017). [Google Scholar]
  • 7.Rossi S, Puglisi A & Benaglia M Additive manufacturing technologies: 3D printing in organic synthesis. ChemCatChem 10, 1512–1525 (2017). [Google Scholar]
  • 8.Parra-Cabrera C, Achille C, Kuhn S & Ameloot R 3D printing in chemical engineering and catalytic technology: structured catalysts, mixers and reactors. Chem. Soc. Rev 47, 209–230 (2018). [DOI] [PubMed] [Google Scholar]
  • 9.Zhou X & Liu C-J Three-dimensional printing for catalytic applications: current status and perspectives. Adv. Funct. Mater 27, 1701134 (2017). [Google Scholar]
  • 10.Baden T et al. Open labware: 3D printing your own lab equipment. PLOS Biol 13, e1002086 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Gordeev EG, Degtyareva ES & Ananikov VP Analysis of 3D printing possibilities for the development of practical applications in synthetic organic chemistry. Russ. Chem. Bull 65, 1637–1643 (2016). [Google Scholar]
  • 12.Kitson PJ et al. 3D printing of versatile reactionware for chemical synthesis. Nat. Protoc 11, 920 (2016). [DOI] [PubMed] [Google Scholar]
  • 13.Lederle F, Meyer F, Kaldun C, Namyslo JC & Hübner EG Sonogashira coupling in 3D-printed NMR cuvettes: synthesis and properties of arylnaphthylalkynes. New J. Chem 41, 1925–1932 (2017). [Google Scholar]
  • 14.Amin R et al. 3D-printed microfluidic devices. Biofabrication 8, 022001 (2016). [DOI] [PubMed] [Google Scholar]
  • 15.Au AK, Huynh W, Horowitz LF & Folch A 3D-printed microfluidics. Angew. Chem. Int. Ed 55, 3862–3881 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bhattacharjee N, Urrios A, Kanga S & Folch A The upcoming 3D-printing revolution in microfluidics. Lab. Chip 16, 1720–1742 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Capel AJ et al. Design and additive manufacture for flow chemistry. Lab. Chip 13, 4583–4590 (2013). [DOI] [PubMed] [Google Scholar]
  • 18.He Y, Wu Y, Fu J-Z, Gao Q & Qiu J-J Developments of 3D printing microfluidics and applications in chemistry and biology: a review. Electroanalysis 28, 1658–1678 (2016). [Google Scholar]
  • 19.Ho CMB, Sum Huan N, Li KHH & Yoon Y-J 3D printed microfluidics for biological applications. Lab. Chip 15, 3627–3637 (2015). [DOI] [PubMed] [Google Scholar]
  • 20.Monaghan T, Harding MJ, Harris RA, Friel RJ & Christie SDR Customisable 3D printed microfluidics for integrated analysis and optimisation. Lab. Chip 16, 3362–3373 (2016). [DOI] [PubMed] [Google Scholar]
  • 21.Tseng P, Murray C, Kim D & Di Carlo D Research highlights: printing the future of microfabrication. Lab. Chip 14, 1491–1495 (2014). [DOI] [PubMed] [Google Scholar]
  • 22.Waldbaur A, Rapp H, Laenge K & Rapp BE Let there be chip-towards rapid prototyping of microfluidic devices: one-step manufacturing processes. Anal. Methods 3, 2681–2716 (2011). [Google Scholar]
  • 23.Saggiomo V & Velders AH Simple 3D printed scaffold-removal method for the fabrication of intricate microfluidic devices. Adv. Sci 2, 1500125 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Dragone V, Sans V, Rosnes MH, Kitson PJ & Cronin L 3D-printed devices for continuous-flow organic chemistry. Beilstein J. Org. Chem 9, 951–959 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Elias Y, von Rohr PR, Bonrath W, Medlock J & Buss A A porous structured reactor for hydrogenation reactions. Chem. Eng. Process 95, 175–185 (2015). [Google Scholar]
  • 26.Nieves-Remacha MJ, Kulkarni AA & Jensen KF Gas-liquid flow and mass transfer in an advanced-flow reactor. Ind. Eng. Chem. Res 52, 8996–9010 (2013). [Google Scholar]
  • 27.Wu KJ, Nappo V & Kuhn S Hydrodynamic study of single- and two-phase flow in an advanced-flow reactor. Ind. Eng. Chem. Res 54, 7554–7564 (2015). [Google Scholar]
  • 28.Kitson PJ et al. Digitization of multi-step organic synthesis in reactionware for on demand pharmaceuticals. Science 359, 314–319 (2018). [DOI] [PubMed] [Google Scholar]
  • 29.Kitson PJ, Marshall RJ, Long DL, Forgan RS & Cronin L 3D printed high-throughput hydrothermal reactionware for discovery, optimization, and scale-up. Angew. Chem. Int. Ed 53, 12723–12728 (2014). [DOI] [PubMed] [Google Scholar]
  • 30.Symes MD et al. Integrated 3D-printed reactionware for chemical synthesis and analysis. Nat. Chem 4, 349–354 (2012). [DOI] [PubMed] [Google Scholar]
  • 31.Kitson PJ, Symes MD, Dragone V & Cronin L Combining 3D printing and liquid handling to produce user-friendly reactionware for chemical synthesis and purification. Chem. Sci 4, 3099–3103 (2013). [Google Scholar]
  • 32.Wang Z, Wang J, Li M, Sun K & Liu C-J Three-dimensional printed acrylonitrile butadiene styrene framework coated with Cu-BTC metal-organic frameworks for the removal of methylene blue. Sci. Rep 4, 5939 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kang H-W et al. A 3D bioprinting system to produce human-scale tissue constructs with structural integrity. Nat. Biotechnol 34, 312 (2016). [DOI] [PubMed] [Google Scholar]
  • 34.Skorski MR, Esenther JM, Ahmed Z, Miller AE & Hartings MR The chemical, mechanical, and physical properties of 3D printed materials composed of TiO2-ABS nanocomposites. Sci. Technol. Adv. Mater 17, 89–97 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Elkoro A & Casanova I 3D printing of structured nanotitania catalysts: a novel binder-free and low-temperature chemical sintering method. 3D Print. Addit. Manuf 5, 220–226 (2018). [Google Scholar]
  • 36.Dul S, Fambri L & Pegoretti A Fused deposition modelling with ABS–graphene nanocomposites. Compos. Part A Appl. Sci. Manuf 85, 181–191 (2016). [Google Scholar]
  • 37.García-Tuñón E et al. Graphene oxide: an all-in-one processing additive for 3D printing. ACS Appl. Mater. Interfaces 9, 32977–32989 (2017). [DOI] [PubMed] [Google Scholar]
  • 38.Zhu C et al. Highly compressible 3D periodic graphene aerogel microlattices. Nat. Commun 6, 6962 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Bible M et al. 3D-printed acrylonitrile butadiene styrene-metal organic framework composite materials and their gas storage properties. 3D Print. Addit. Manuf 5, 63–72 (2018). [Google Scholar]
  • 40.Evans KA et al. Chemically active, porous 3D-printed thermoplastic composites. ACS Appl. Mater. Interfaces 10, 15112–15121 (2018). [DOI] [PubMed] [Google Scholar]
  • 41.Kreider MC et al. Toward 3D printed hydrogen storage materials made with ABS-MOF composites. Polym. Adv. Technol 29, 867–873 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Farahani RD & Dube M Printing polymer nanocomposites and composites in three dimensions. Adv. Eng. Mater 20, 1700539 (2018). [Google Scholar]
  • 43.Ligon SC, Liska R, Stampfl J, Gurr M & Mulhaupt R Polymers for 3D printing and customized additive manufacturing. Chem. Rev 117, 10212–10290 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Olivera S, Muralidhara HB, Venkatesh K, Gopalakrishna K & Vivek CS Plating on acrylonitrile-butadiene-styrene (ABS) plastic: a review. J. Mater. Sci 51, 3657–3674 (2016). [Google Scholar]
  • 45.Kacar E et al. Functionalized hybrid coatings on ABS surfaces by PLD and dip coatings. J. Inorg. Organomet. Polym. Mater 26, 895–906 (2016). [Google Scholar]
  • 46.Kukushkin VY & Pombeiro AJL Metal-mediated and metal-catalyzed hydrolysis of nitriles. Inorg. Chim. Acta 358, 1–21 (2005). [Google Scholar]
  • 47.Makadia HK & Siegel SJ Poly lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 3, 1377–1397 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Peterson GI, Larsen MB, Ganter MA, Storti DW & Boydston AJ 3D-printed mechanochromic materials. ACS Appl. Mater. Interfaces 7, 577–583 (2015). [DOI] [PubMed] [Google Scholar]
  • 49.Lars K, Anton B, Aldrik V & Vittorio S Gold nanoparticles embedded in a polymer as a 3D-printable dichroic nanocomposite material. Beilstein J. Nanotechnol 10, 442–447 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Shi Z et al. Renewable metal–organic-frameworks-coated 3D printing film for removal of malachite green. RSC Adv 7, 49947–49952 (2017). [Google Scholar]
  • 51.Cameron NR & Barbetta A The influence of porogen type on the porosity, surface area and morphology of poly(divinylbenzene) PolyHIPE foams. J. Mater. Chem 10, 2466–2471 (2000). [Google Scholar]
  • 52.Lewis JA Direct ink writing of 3D functional materials. Adv. Funct. Mater 16, 2193–2204 (2006). [Google Scholar]
  • 53.Kolesky DB et al. 3D bioprinting of vascularized, heterogeneous cell-laden tissue constructs. Adv. Mater 26, 3124–3130 (2014). [DOI] [PubMed] [Google Scholar]
  • 54.Hinton TJ et al. Three-dimensional printing of complex biological structures by freeform reversible embedding of suspended hydrogels. Sci. Adv 1, e1500758 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Weng B, Shepherd RL, Crowley K, Killard AJ & Wallace GG Printing conducting polymers. Analyst 135, 2779–2789 (2010). [DOI] [PubMed] [Google Scholar]
  • 56.Holness FB & Aaron DP Direct ink writing of 3D conductive polyaniline structures and rheological modelling. Smart Mater. Struct 27, 015006 (2018). [Google Scholar]
  • 57.Wang Z et al. Three-dimensional printing of polyaniline/reduced graphene oxide composite for high-performance planar supercapacitor. ACS Appl. Mater. Interfaces 10, 10437–10444 (2018). [DOI] [PubMed] [Google Scholar]
  • 58.Wei Q et al. Printable hybrid hydrogel by dual enzymatic polymerization with superactivity. Chem. Sci 7, 2748–2752 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Stuecker JN, Miller JE, Ferrizz RE, Mudd JE & Cesarano J Advanced support structures for enhanced catalytic activity. Ind. Eng. Chem. Res 43, 51–55 (2004). [Google Scholar]
  • 60.Noyen J, Wilde A, Schroeven M, Mullens S & Luyten J Ceramic processing techniques for catalyst design: formation, properties, and catalytic example of ZSM-5 on 3-dimensional fiber deposition support structures. Int. J. Appl. Ceram. Technol 9, 902–910 (2012). [Google Scholar]
  • 61.Kong YL et al. 3D printed quantum dot light-emitting diodes. Nano Lett 14, 7017–7023 (2014). [DOI] [PubMed] [Google Scholar]
  • 62.Tubío CR et al. 3D printing of a heterogeneous copper-based catalyst. J. Catalysis 334, 110–115 (2016). [Google Scholar]
  • 63.Zhu C et al. Toward digitally controlled catalyst architectures: hierarchical nanoporous gold via 3D printing. Sci. Adv 4, eaas9459 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Lefevere J, Gysen M, Mullens S, Meynen V & Van Noyen J The benefit of design of support architectures for zeolite coated structured catalysts for methanol-to-olefin conversion. Catal. Today 216, 18–23 (2013). [Google Scholar]
  • 65.Thakkar H, Eastman S, Al-Naddaf Q, Rownaghi AA & Rezaei F 3D-printed metal–organic framework monoliths for gas adsorption processes. ACS Appl. Mater. Interfaces 9, 35908–35916 (2017). [DOI] [PubMed] [Google Scholar]
  • 66.Goyos-Ball L et al. Mechanical and biological evaluation of 3D printed 10CeTZP-Al2O3 structures. J. Eur. Ceram. Soc 37, 3151–3158 (2017). [Google Scholar]
  • 67.Billiet T, Gevaert E, De Schryver T, Cornelissen M & Dubruel P The 3D printing of gelatin methacrylamide cell-laden tissue-engineered constructs with high cell viability. Biomaterials 35, 49–62 (2014). [DOI] [PubMed] [Google Scholar]
  • 68.Jia W et al. Direct 3D bioprinting of perfusable vascular constructs using a blend bioink. Biomaterials 106, 58–68 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Rambo CR, Travitzky N, Zimmermann K & Greil P Synthesis of TiC/Ti–Cu composites by pressureless reactive infiltration of TiCu alloy into carbon preforms fabricated by 3D-printing. Mater. Lett 59, 1028–1031 (2005). [Google Scholar]
  • 70.Lam CXF, Mo XM, Teoh SH & Hutmacher DW Scaffold development using 3D printing with a starch-based polymer. Mater. Sci. Eng. C 20, 49–56 (2002). [Google Scholar]
  • 71.Motealleh A et al. Understanding the role of dip-coating process parameters in the mechanical performance of polymer-coated bioglass robocast scaffolds. J. Mechan. Behav. Biomed. Mater 64, 253–261 (2016). [DOI] [PubMed] [Google Scholar]
  • 72.Díaz-Marta AS et al. Three-dimensional printing in catalysis: combining 3D heterogeneous copper and palladium catalysts for multicatalytic multicomponent reactions. ACS Catal 8, 392–404 (2018). [Google Scholar]
  • 73.Liao HT, Lee MY, Tsai WW, Wang HC & Lu WC Osteogenesis of adipose-derived stem cells on polycaprolactone-beta-tricalcium phosphate scaffold fabricated via selective laser sintering and surface coating with collagen type I. J. Tissue Eng. Regen. Med 10, E337–E353 (2016). [DOI] [PubMed] [Google Scholar]
  • 74.Duan B & Wang M Customized Ca-P/PHBV nanocomposite scaffolds for bone tissue engineering: design, fabrication, surface modification and sustained release of growth factor. J. R. Soc. Interface 7, S615–S629 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Alaman J, Alicante R, Pena JI & Sanchez-Somolinos C Inkjet printing of functional materials for optical and photonic applications. Materials 9, 910 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Katz JE, Gingrich TR, Santori EA & Lewis NS Combinatorial synthesis and high-throughput photopotential and photocurrent screening of mixed-metal oxides for photoelectrochemical water splitting. Energy Environ. Sci 2, 103–112 (2009). [Google Scholar]
  • 77.Liu X et al. Inkjet printing assisted synthesis of multicomponent mesoporous metal oxides for ultrafast catalyst exploration. Nano Lett 12, 5733–5739 (2012). [DOI] [PubMed] [Google Scholar]
  • 78.Seley D, Ayers K & Parkinson BA Combinatorial search for improved metal oxide oxygen evolution electrocatalysts in acidic electrolytes. ACS Comb. Sci 15, 82–89 (2013). [DOI] [PubMed] [Google Scholar]
  • 79.Melchels FPW, Feijen J & Grijpma DW A review on stereolithography and its applications in biomedical engineering. Biomaterials 31, 6121–6130 (2010). [DOI] [PubMed] [Google Scholar]
  • 80.Zheng X et al. Ultralight, ultrastiff mechanical metamaterials. Science 344, 1373 (2014). [DOI] [PubMed] [Google Scholar]
  • 81.Hensleigh RM et al. Additive manufacturing of complex micro-architected graphene aerogels. Mater. Horizons 5, 1035–1041 (2018). [Google Scholar]
  • 82.Mu Q et al. Digital light processing 3D printing of conductive complex structures. Addit. Manuf 18, 74–83 (2017). [Google Scholar]
  • 83.Xing J-F, Zheng M-L & Duan X-M Two-photon polymerization microfabrication of hydrogels: an advanced 3D printing technology for tissue engineering and drug delivery. Chem. Soc. Rev 44, 5031–5039 (2015). [DOI] [PubMed] [Google Scholar]
  • 84.Tumbleston JR et al. Continuous liquid interface production of 3D objects. Science 347, 1349 (2015). [DOI] [PubMed] [Google Scholar]
  • 85.Wang X et al. Initiator-integrated 3D printing enables the formation of complex metallic architectures. ACS Appl. Mater. Interfaces 6, 2583–2587 (2014). [DOI] [PubMed] [Google Scholar]
  • 86.Yue J et al. 3D-printable antimicrobial composite resins. Adv. Funct. Mater 25, 6756–6767 (2015). [Google Scholar]
  • 87.Manzano JS, Weinstein ZB, Sadow AD & Slowing II Direct 3D printing of catalytically active structures. ACS Catal 7, 7567–7577 (2017). [Google Scholar]
  • 88.Thrasher CJ, Schwartz JJ & Boydston AJ Modular elastomer photoresins for digital light processing additive manufacturing. ACS Appl. Mater. Interfaces 9, 39708–39716 (2017). [DOI] [PubMed] [Google Scholar]
  • 89.Zhu W et al. A 3D printed nano bone matrix for characterization of breast cancer cell and osteoblast interactions. Nanotechnology 27, 315103 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Nowicki M, Castro NJ, Rao R, Plesniak M & Zhang LJG Integrating three-dimensional printing and nanotechnology for musculoskeletal regeneration. Nanotechnology 28, 382001 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Gao G et al. Bioactive nanoparticles stimulate bone tissue formation in bioprinted three-dimensional scaffold and human mesenchymal stem cells. Biotechnol. J 9, 1304–1311 (2014). [DOI] [PubMed] [Google Scholar]
  • 92.Cheng F et al. Magnetic control of MOF crystal orientation and alignment. Chemistry 23, 15578–15582 (2017). [DOI] [PubMed] [Google Scholar]
  • 93.Halevi O, Tan JMR, Lee PS & Magdassi S Hydrolytically stable MOF in 3D-printed structures. Adv. Sustain. Syst 2, 1700150 (2018). [Google Scholar]
  • 94.Seyed Farid Seyed S et al. A review on powder-based additive manufacturing for tissue engineering: selective laser sintering and inkjet 3D printing. Sci. Technol. Adv. Mater 16, 033502 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Avril A et al. Continuous flow hydrogenations using novel catalytic static mixers inside a tubular reactor. React. Chem. Eng 2, 180–188 (2017). [Google Scholar]
  • 96.Kim SS et al. Survival and function of hepatocytes on a novel three-dimensional synthetic biodegradable polymer scaffold with an intrinsic network of channels. Ann. Surg 228, 8–13 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Wu BM et al. Solid free-form fabrication of drug delivery devices. J. Control. Release 40, 77–87 (1996). [Google Scholar]
  • 98.Seitz H, Rieder W, Irsen S, Leukers B & Tille C Three-dimensional printing of porous ceramic scaffolds for bone tissue engineering. J. Biomed. Mater. Res. B 74, 782–788 (2005). [DOI] [PubMed] [Google Scholar]
  • 99.Chisholm G, Kitson PJ, Kirkaldy ND, Bloor LG & Cronin L 3D printed flow plates for the electrolysis of water: an economic and adaptable approach to device manufacture. Energy Environ. Sci 7, 3026–3032 (2014). [Google Scholar]
  • 100.Michorczyk P, Hedrzak E & Wegrzyniak A Preparation of monolithic catalysts using 3D printed templates for oxidative coupling of methane. J. Mater. Chem. A 4, 18753–18756 (2016). [Google Scholar]
  • 101.Mannoor MS et al. 3D printed bionic ears. Nano Lett 13, 2634–2639 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Ni Y et al. A review of 3D-printed sensors. Appl. Spectrosc. Rev 52, 623–652 (2017). [Google Scholar]
  • 103.Jackson JA et al. Field responsive mechanical metamaterials. Sci. Adv 4, eaau6419 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Weidenbach M et al. 3D printed dielectric rectangular waveguides, splitters and couplers for 120 GHz. Opt. Express 24, 28968–28976 (2016). [DOI] [PubMed] [Google Scholar]
  • 105.Ambrosi A & Pumera M 3D-printing technologies for electrochemical applications. Chem. Soc. Rev 45, 2740–2755 (2016). [DOI] [PubMed] [Google Scholar]
  • 106.Wei X et al. 3D printable graphene composite. Sci. Rep 5, 11181–11181 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Huang L et al. Ultrafast digital printing toward 4D shape changing materials. Adv. Mater 29, 1605390 (2016). [DOI] [PubMed] [Google Scholar]
  • 108.Sydney Gladman A, Matsumoto EA, Nuzzo RG, Mahadevan L & Lewis JA Biomimetic 4D printing. Nat. Mater 15, 413 (2016). [DOI] [PubMed] [Google Scholar]
  • 109.Ge Q et al. Multimaterial 4D printing with tailorable shape memory polymers. Sci. Rep 6, 31110 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Diercks CS, Kalmutzki MJ, Diercks NJ & Yaghi OM Conceptual advances from Werner complexes to metal–organic frameworks. ACS Cent. Sci 4, 1457–1464 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Wales DJ et al. 3D-printable photochromic molecular materials for reversible information storage. Adv. Mater 30, 1800159 (2018). [DOI] [PubMed] [Google Scholar]
  • 112.Yao B et al. Efficient 3D printed pseudocapacitive electrodes with ultrahigh MnO2 loading. Joule 3, 459–470 (2018). [Google Scholar]
  • 113.Zhu C et al. 3D printed functional nanomaterials for electrochemical energy storage. Nano Today 15, 107–120 (2017). [Google Scholar]
  • 114.Donderwinkel I, van Hest JCM & Cameron NR Bio-inks for 3D bioprinting: recent advances and future prospects. Polym. Chem 8, 4451–4471 (2017). [Google Scholar]
  • 115.Hollister SJ Porous scaffold design for tissue engineering. Nat. Mater 4, 518–524 (2005). [DOI] [PubMed] [Google Scholar]
  • 116.Lu P, Takai K, Weaver VM & Werb Z Extracellular matrix degradation and remodeling in development and disease. Cold Spring Harb. Perspect. Biol 3, a005058 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Levato R et al. Biofabrication of tissue constructs by 3D bioprinting of cell-laden microcarriers. Biofabrication 6, 035020 (2014). [DOI] [PubMed] [Google Scholar]
  • 118.Cui H et al. Hierarchical fabrication of engineered vascularized bone biphasic constructs via dual 3D bioprinting: integrating regional bioactive factors into architectural design. Adv. Healthc. Mater 5, 2174–2181 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Holmes B, Bulusu K, Plesniak M & Zhang LG A synergistic approach to the design, fabrication and evaluation of 3D printed micro and nano featured scaffolds for vascularized bone tissue repair. Nanotechnology 27, 064001 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Zhou X & Liu C-J Three-dimensional printing of porous carbon structures with tailorable pore sizes. Catal. Today 10.1016/j.cattod.2018.05.044 (2018). [DOI] [Google Scholar]

RESOURCES