In this study, we explore the integration of additive manufacturing with the traditional lost wax investment casting process to produce high-quality impeller components. The lost wax investment casting method is widely recognized for its ability to create intricate and precise metal parts, particularly in industries such as aerospace and industrial gas turbines. By combining this with fused deposition modeling (FDM) 3D printing, we aim to streamline the production cycle, reduce costs, and minimize defects associated with conventional mold-making techniques. Our focus is on designing and optimizing the gating system for an impeller, utilizing numerical simulation software to predict and address potential issues before physical prototyping. This approach not only enhances the efficiency of the lost wax investment casting process but also leverages the flexibility of additive manufacturing to overcome limitations in complex geometries.
The lost wax investment casting process involves several critical steps: creating a sacrificial pattern, coating it with ceramic slurry to form a mold, removing the pattern through heating, and pouring molten metal into the preheated mold. After solidification, the mold is broken away, and the casting is finished by removing gates and performing surface treatments. Traditionally, wax patterns are used, but we substitute these with 3D-printed polylactic acid (PLA) models. This substitution allows for rapid iteration and customization, which is essential for components like impellers that feature complex, curved blades. Our investigation centers on an impeller with a base diameter of 104 mm, top diameter of 25 mm, and side width of 49 mm, as detailed in the initial design. The blade wall thickness is set at 2.5 mm, based on standard lost wax investment casting guidelines for such dimensions.
To ensure the success of the lost wax investment casting process, we designed three distinct gating systems: top-gating, side-gating, and bottom-gating. Each system was evaluated using AnyCasting software to simulate the filling and solidification phases. The simulations accounted for key parameters, including pouring temperature, mold preheat temperature, and pouring velocity, to identify the optimal configuration that minimizes defects like turbulence, shrinkage, and porosity. Through this numerical analysis, we selected the bottom-gating system for experimental validation, as it demonstrated superior performance in terms of sequential solidification and defect distribution. Subsequently, we fabricated the pattern via FDM 3D printing and conducted actual lost wax investment casting trials with ZL104 aluminum alloy, achieving a defect-free impeller casting after post-processing.
The impeller’s geometry, as illustrated in the design phase, presents challenges due to its curved blades and thin sections, which can lead to unstable metal flow during pouring. In lost wax investment casting, such complexities often result in surface imperfections and internal defects if not properly managed. Our use of additive manufacturing for pattern creation addresses this by enabling precise control over the pattern’s dimensions, thereby reducing the likelihood of deviations in the final casting. The following sections delve into the specifics of our methodology, including the gating system designs, simulation results, and experimental procedures, all emphasizing the synergies between additive manufacturing and lost wax investment casting.
Structural Analysis of the Impeller
The impeller under investigation has a complex structure characterized by multiple curved blades radiating from a central hub. Key dimensions include a bottom diameter of 104 mm, top diameter of 25 mm, and a side width of 49 mm, with blade thickness maintained at 2.5 mm to ensure structural integrity during the lost wax investment casting process. This geometry was modeled in 3D CAD software to facilitate both simulation and additive manufacturing. The blades’ curvature and the overall asymmetry necessitate a careful approach to gating design, as improper metal flow can cause turbulence, leading to defects such as misruns and cold shuts. In lost wax investment casting, the pattern’s accuracy is paramount; thus, we employed FDM 3D printing with PLA material to produce patterns that faithfully replicate the digital model. The table below summarizes the critical dimensions and their implications for the lost wax investment casting process.
| Parameter | Value (mm) | Significance in Lost Wax Investment Casting | 
|---|---|---|
| Base Diameter | 104 | Determines mold size and thermal dynamics during pouring | 
| Top Diameter | 25 | Influences metal flow velocity and solidification patterns | 
| Side Width | 49 | Affects blade stability and defect susceptibility | 
| Blade Thickness | 2.5 | Critical for avoiding shrinkage and ensuring complete filling | 
In lost wax investment casting, the mold material plays a vital role in heat transfer and surface finish. We selected a high-temperature gypsum-based ceramic for the shell due to its low thermal conductivity and high replication accuracy, which helps maintain dimensional stability during the high-temperature baking phase. The relationship between mold properties and casting quality can be expressed through the heat transfer coefficient, which influences the solidification rate. For instance, the rate of heat loss in the mold can be modeled using Fourier’s law of heat conduction:
$$ q = -k \nabla T $$
where \( q \) is the heat flux, \( k \) is the thermal conductivity of the mold, and \( \nabla T \) is the temperature gradient. In lost wax investment casting, a lower \( k \) value, as in gypsum, reduces thermal shocks and minimizes defects like hot tearing. Additionally, the pattern removal temperature must be optimized to prevent residual stresses; we used a step-wise heating cycle up to 750°C to ensure complete PLA degradation without damaging the ceramic shell.
Gating System Design and Numerical Simulation
We designed three gating systems for the impeller: top-gating, side-gating, and bottom-gating, each with unique flow characteristics. The primary goal was to achieve laminar flow and directional solidification, which are essential for high-quality lost wax investment casting. Numerical simulations using AnyCasting software provided insights into filling patterns, temperature distributions, and defect formations. The software solves the Navier-Stokes equations for fluid flow and energy equations for heat transfer, allowing us to predict potential issues. The general form of the continuity and momentum equations for incompressible flow during pouring is:
$$ \nabla \cdot \mathbf{v} = 0 $$
$$ \rho \left( \frac{\partial \mathbf{v}}{\partial t} + \mathbf{v} \cdot \nabla \mathbf{v} \right) = -\nabla p + \mu \nabla^2 \mathbf{v} + \rho \mathbf{g} $$
where \( \mathbf{v} \) is the velocity vector, \( \rho \) is the density of ZL104 aluminum alloy, \( p \) is pressure, \( \mu \) is dynamic viscosity, and \( \mathbf{g} \) is gravitational acceleration. For lost wax investment casting, these equations help simulate how metal enters the mold cavity, with parameters adjusted based on the gating design.
The simulation parameters were consistent across all designs, as outlined in the table below, to ensure fair comparison. Pouring temperature was set at 750°C, mold preheat at 650°C, and pouring velocity at 25 cm/s, reflecting typical conditions for aluminum lost wax investment casting.
| Parameter | Value | Unit | 
|---|---|---|
| Pouring Temperature | 750 | °C | 
| Mold Preheat Temperature | 650 | °C | 
| Pouring Velocity | 25 | cm/s | 
| Alloy Material | ZL104 | – | 
| Mold Material | Gypsum-based Ceramic | – | 
For the top-gating system, simulations revealed turbulent flow as metal entered from the top, causing splashing and potential gas entrapment. The filling time was approximately 8.5 seconds, with defects concentrated at the base due to inverse temperature gradients. The defect indicator, combining temperature gradient and inverse interface velocity, highlighted areas prone to shrinkage, calculated as:
$$ D = \left( \frac{\partial T}{\partial x} \right) \times \left( \frac{1}{v} \right) $$
where \( D \) is the defect score, \( \frac{\partial T}{\partial x} \) is the temperature gradient along the solidification direction, and \( v \) is the interface velocity. In top-gating, high \( D \) values correlated with shrinkage porosity in the lower regions.
Side-gating showed improved flow stability, with a filling time of 13 seconds. However, rapid filling led to minor air entrapment near the gate-blade junctions, resulting in localized porosity. The solidification sequence was more uniform, but defect predictions indicated susceptibility in the gating areas. Bottom-gating, with a filling time of 15 seconds, demonstrated the best performance, with sequential solidification from the bottom upward and minimal defects on the blades. The shrinkage porosity score was predominantly confined to the gate corners, affirming its suitability for lost wax investment casting of complex impellers.
Based on these simulations, we proceeded with the bottom-gating design for experimental validation. The 3D-printed pattern for this system was fabricated using FDM technology, ensuring precise replication of the gating geometry. This step underscores the advantage of combining additive manufacturing with lost wax investment casting, as it allows for rapid prototyping and iterative design improvements without the need for expensive tooling.

Experimental Procedure in Lost Wax Investment Casting
Our experimental approach followed the standard lost wax investment casting protocol, adapted for 3D-printed patterns. We began by preparing the PLA pattern of the impeller with the bottom-gating system, printed at a layer resolution of 0.2 mm to achieve smooth surfaces. The pattern was then coated with a ceramic slurry composed of gypsum and water in a 100:45 ratio, carefully stirred to eliminate air bubbles that could cause defects in the final casting. This slurry application is critical in lost wax investment casting, as it forms the mold cavity upon pattern removal. After coating, the pattern was placed in a steel flask, and additional slurry was poured to encapsulate it fully, ensuring even distribution around the blades.
The mold was allowed to harden at room temperature before undergoing a controlled baking process in a resistance furnace. The temperature was ramped up to 400°C for 15 minutes, then to 600°C for 30 minutes, and finally to 750°C for 2 hours to completely vaporize the PLA pattern. This gradual heating prevents thermal shock and ensures a clean mold cavity, which is essential for achieving high surface quality in lost wax investment casting. The mold was then cooled to 600°C before pouring to maintain an optimal temperature gradient for metal solidification.
For metal pouring, ZL104 aluminum alloy was melted in a crucible and poured into the preheated mold at 750°C. The pouring process was monitored to ensure a steady flow, minimizing turbulence. After solidification, the mold was broken away, and the casting was extracted, followed by removal of the gating system and surface finishing using sandpaper. The resulting impeller exhibited no visible defects such as porosity or cold shuts, demonstrating the effectiveness of the bottom-gating design in lost wax investment casting. The table below summarizes the key steps and parameters in our experimental lost wax investment casting process.
| Step | Description | Parameters | 
|---|---|---|
| Pattern Printing | FDM 3D printing with PLA | Layer height: 0.2 mm | 
| Slurry Preparation | Gypsum-water mixture | Ratio: 100:45 by weight | 
| Mold Baking | Pattern removal and mold preheat | Max temperature: 750°C, hold time: 2 h | 
| Metal Pouring | ZL104 aluminum alloy | Pouring temperature: 750°C | 
| Post-Processing | Gate removal and surface finishing | Manual grinding with sandpaper | 
The success of this lost wax investment casting trial highlights the role of additive manufacturing in reducing lead times and costs. By integrating 3D printing, we eliminated the need for wax pattern fabrication, which is often time-consuming and prone to errors in complex geometries. Moreover, the use of numerical simulation allowed us to preemptively address potential defects, reinforcing the reliability of lost wax investment casting for precision components.
Results and Discussion
The final impeller casting produced via lost wax investment casting met all dimensional and quality specifications, with no significant defects observed on the blade surfaces. Numerical simulations accurately predicted the defect locations, particularly in the gating areas, which aligned with the experimental outcomes. For instance, the shrinkage porosity fraction in the bottom-gating system was negligible on the blades, confirming the simulation’s validity. This correlation underscores the importance of CAE tools in optimizing lost wax investment casting processes for additive manufacturing applications.
In lost wax investment casting, the solidification behavior directly impacts mechanical properties. We analyzed the cooling curve using the Fourier number for transient heat conduction, which relates the time scale of solidification to the mold’s thermal properties:
$$ Fo = \frac{\alpha t}{L^2} $$
where \( Fo \) is the Fourier number, \( \alpha \) is the thermal diffusivity of the mold, \( t \) is time, and \( L \) is the characteristic length of the casting. For our impeller, a higher \( Fo \) value indicated slower cooling in thicker sections, which helped avoid thermal stresses. The bottom-gating system promoted a favorable temperature gradient, reducing the risk of hot tears—a common issue in lost wax investment casting of thin-walled components.
Furthermore, the combination of additive manufacturing and lost wax investment casting enabled rapid iteration; had the simulations indicated major flaws, we could have redesigned the gating system and reprinted the pattern with minimal delay. This flexibility is particularly beneficial for low-volume production or prototyping, where traditional lost wax investment casting might be prohibitive due to tooling costs. Our findings suggest that this integrated approach can be extended to other complex parts, such as turbine blades or medical implants, where precision and efficiency are paramount.
Despite the successes, challenges remain in lost wax investment casting with 3D-printed patterns, such as ensuring complete pattern removal without residue and managing the thermal expansion mismatch between the pattern and ceramic shell. Future work could focus on optimizing the baking cycle and exploring alternative pattern materials to enhance the lost wax investment casting process further.
Conclusion
In this research, we successfully demonstrated the integration of additive manufacturing with lost wax investment casting to produce a high-quality aluminum impeller. Through numerical simulation and experimental validation, we identified the bottom-gating system as the optimal design, minimizing defects and ensuring uniform solidification. The use of FDM 3D printing for pattern fabrication significantly reduced the time and cost associated with traditional lost wax investment casting, while AnyCasting software provided accurate predictions of filling and solidification behavior. Our results confirm that lost wax investment casting, when combined with modern technologies, can achieve superior outcomes for complex geometries, paving the way for broader applications in precision engineering. This study underscores the transformative potential of additive manufacturing in enhancing lost wax investment casting processes, offering a robust solution for industries requiring intricate metal components.
