Categories
Misc

Developers Look to OpenUSD in Era of AI and Industrial Digitalization

From smart factories to next-generation railway systems, developers and enterprises across the world are racing to fuel industrial digitalization opportunities at every scale. Key to this is the open-source Universal Scene Description (USD) framework, or OpenUSD, along with metaverse applications powered by AI. OpenUSD, originally developed by Pixar for large-scale feature film pipelines for animation Read article >

Categories
Misc

Sensing New Frontiers with Neural Lidar Fields for Autonomous Vehicle Simulation

A still view of a lidar point cloud in a driving scene generated by neural lidar fields.Autonomous vehicle (AV) development requires massive amounts of sensor data for perception development. Developers typically get this data from two…A still view of a lidar point cloud in a driving scene generated by neural lidar fields.

Autonomous vehicle (AV) development requires massive amounts of sensor data for perception development.

Developers typically get this data from two sources—replay streams of real-world drives or simulation. However, real-world datasets offer limited flexibility, as the data is fixed to only the objects, events, and view angles captured by the physical sensors. It is also difficult to simulate the detail and imperfection of real-world conditions—such as sensor noise or occlusions—at scale.

Neural fields have gained significant traction in recent years. These AI tools capture real-world content and simulate it from novel viewpoints with high levels of realism, achieving the fidelity and diversity required for AV simulation.

At NVIDIA GTC 2022, we showed how we use neural reconstruction to build a 3D scene from recorded camera sensor data in simulation, which can then be rendered from novel views. A paper we published for ICCV 2023—which runs Oct. 2 to Oct.6—details how we applied a similar approach to address these challenges in synthesizing lidar data.

GIF showing 360-degree lidar point cloud returns of a driving scene with cars and free space.
Figure 1. An example novel viewpoint rendered by neural lidar fields

The method, called neural lidar fields, optimizes a neural radiance field (NeRF)-like representation from lidar measurements that enables synthesizing realistic lidar scans from entirely new viewpoints. It combines neural rendering with a physically based lidar model to accurately reproduce sensor behaviors—such as beam divergence, secondary returns, and ray dropping.

With neural lidar fields, we can achieve improved realism of novel views, narrowing the domain gap with real lidar data recordings. In doing so, we can improve the scalability of lidar sensor simulation and accelerate AV development.

By applying neural rendering techniques such as neural lidar fields in NVIDIA Omniverse, AV developers can bypass the time– and cost-intensive process of rebuilding real-world scenes by hand. They can bring physical sensors into a scalable and repeatable simulation.

Novel view synthesis

While replaying recorded data is a key component of testing and validation, it is critical to also simulate new scenarios for the AV system to experience.

These scenarios make it possible to test situations where the vehicle deviates from the original trajectory. It will view the world from novel views. This benefit also extends to testing a sensor suite on a different vehicle type, where the rig may be positioned differently (for example, switching from a sedan to an SUV).

With the ability to modify sensor properties such as beam divergence and ray pattern, we can also use a different lidar type in simulation than the sensor that originally recorded the data.

However, previous explicit approaches to simulating novel views have proven cumbersome and often inaccurate. First, surface representation—such as surfels or a triangular mesh—must be extracted from scanned lidar point clouds. Then, lidar measurements are simulated from a novel viewpoint by casting rays and intersecting them with the surface model.

These methods—known as explicit reconstruction—introduce noticeable errors in the rendering as well as assuming a perfect lidar model with no divergence of beams.

Neural lidar fields method

Rather than rely on an error-prone reconstruction pipeline, the neural lidar fields method takes a NeRF-style approach. It is based on neural scene representation and sensor modeling, which is directly optimized to render sensor measurements. This results in a more realistic output.

Specifically, we used an improved, lidar-specific volume rendering procedure, which creates range and intensity measurements from the 3D scene. Then, we added beam divergence for improved realism. We took into account that lidar works as an active sensor—rather than a passive one like a camera. This, along with characteristics such as beam divergence, enabled us to reproduce sensor properties, including dropped rays and multiple returns.

To test the accuracy of the neural lidar fields, we ran the scenes in a lidar simulator, comparing results with a variety of viewpoints taken at different distances from the original scan.

These scans were then compared with real data from the Waymo Open dataset, using metrics such as real-world intensities, ray drop, and secondary returns to evaluate fidelity. We also used real data to validate the accuracy of the neural lidar fields’ view synthesis in challenging scenes.

A series of line graphs showing peaks in radiance, density, and weight, where the neural lidar field accurately models the real lidar beam.
Figure 2. Neural lidar fields model the waveform

In Figure 2, the neural lidar fields accurately reproduce the waveform properties. The top row shows that the first surface fully scatters the lidar energy. The other rows shows that neural lidar fields estimate range through peak detection on the computed weights followed by volume rendering-based range refinement.

Results

Using these evaluation methods, we compared neural lidar field-synthesized lidar views with traditional reconstruction processes.

By accounting for real-world lidar characteristics, neural lidar field views reduced range errors and improved performance compared with explicit reconstruction. We also found the implicit method synthesized challenging scenes with high accuracy.

After we established performance, we then validated the neural lidar field-generated scans using two low-level perception tasks: point cloud registration and semantic segmentation.

We applied the same model to both real-world lidar scans and various synthesized scans to evaluate how well the scans maintained accuracy. We found that neural lidar fields outperformed the baseline methods on datasets with complex geometry and high noise levels.

Comparisons of three lidar scans: ground truth, neural lidar fields, and LidarSim, showing the neural lidar fields accurately reflecting the same scan as the ground truth scene.
Figure 3. Qualitative visualization of lidar novel view synthesis on the Waymo dataset.

For semantic segmentation, we applied the same pretrained model to both real and synthetic lidar scans. Neural lidar fields achieved the highest recall for vehicles, which are especially difficult to render due to sensor noise such as dual returns and ray drops.

While neural lidar fields are still an active research method, it is a critical tool for scalable AV simulation. Next, we plan to focus on generalizing the networks across scenes and handling dynamic environments. Eventually, developers on Omniverse and the NVIDIA DRIVE Sim AV simulator will be able to tap into these AI-powered approaches for accelerated and physically based simulation. 

For more information about neural lidar fields and our development and evaluation methods, see the Neural LiDAR Fields for Novel View Synthesis paper.

Acknowledgments

We would like to thank our collaborators at ETH Zurich, Shengyu Huang and Konrad Schindler, as well as Zan Gojcic, Zian Wang, Francis Williams, Yoni Kasten, Sanja Fidler, and Or Litany from the NVIDIA Research team.

Categories
Misc

Modeling Earth’s Atmosphere with Spherical Fourier Neural Operators

Machine learning-based weather prediction has emerged as a promising complement to traditional numerical weather prediction (NWP) models. Models such as NVIDIA…

Machine learning-based weather prediction has emerged as a promising complement to traditional numerical weather prediction (NWP) models. Models such as NVIDIA FourCastNet have demonstrated that the computational time for generating weather forecasts can be reduced from hours to mere seconds, a significant improvement to current NWP-based workflows.

Traditional methods are formulated from first principles and typically require a timestep restriction to guarantee the accuracy of the underlying numerical method. ML-based approaches do not come with such restrictions, and their uniform memory access patterns are ideally suited for GPUs.

However, these methods are purely data-driven, and you may rightfully ask:

  • How can we trust these models?
  • How well do they generalize?
  • How can we further increase their skill, trustworthiness, and explainability, if they are not formulated from first principles?

In this post, we discuss spherical Fourier neural operators (SFNOs), physical systems on the sphere, the importance of symmetries, and how SFNOs are implemented using the spherical harmonic transform (SHT). For more information about the math, see the ICML paper, Spherical Fourier Neural Operators: Learning Stable Dynamics on the Sphere.

The importance of symmetries

A video depicting two spherical plots, comparing a 5 month roll-out using SFNO to ground-truth data.
Figure 1: 5-month-long rollout of SFNO. Surface windspeed predictions with SFNO and ground truth data are compared to each other.

A potential approach to creating principled and trustworthy models involves formulating them in a manner akin to the formulation of physical laws.

Physical laws are typically formulated from symmetry considerations:

  • We do not expect physics to depend on the frame of reference.
  • We further expect underlying physical laws to remain unchanged if the frame of reference is altered.

In the context of physical systems on the sphere, changes in the frame of reference are accomplished through rotations. Thus, we strive to establish a formulation that remains equivariant under rotations.

Current ML-based weather prediction models treat the state of the atmosphere as a discrete series of vectors representing physical quantities of interest at various spatial locations over time. Any of these vectors are updated by a learned function, which maps the current state to the next state in the sequence.

In plain terms, we ask a neural network to consecutively predict the weather in the next time step when showing it the weather of today. This is comparable to the integration of a physical system using traditional methods, with the caveat of having learned the dynamics in a purely data-driven manner, as opposed to deriving them from physical laws. This approach enables significantly larger time steps as opposed to traditional methods.

The task at hand can thus be understood as learning image-to-image mappings between finite-dimensional vector spaces.

While a broad variety of neural network topologies such as U-Nets are applicable to this task, such approaches ignore the functional nature of the problem. Both input and output are functions and their evolution is governed by partial differential equations.

Traditional ML approaches such as U-Nets ignore this, as they learn a map at a fixed resolution. Neural operators generalize neural networks to solve this problem. Rather than learning maps between finite-dimensional spaces, they learn an operator that can directly map one function to another.

As such, Fourier neural operators (FNOs) provide a powerful framework for learning maps between function spaces and approximating the solution operator of PDEs, which maps one state to the next.

However, classical FNOs are defined in Cartesian space, whose associated symmetries differ from those of the sphere. In practice, ignoring the geometry and pretending that Earth is a periodic rectangle leads to artifacts, which accumulate on long rollouts, due to the autoregressive nature of the model. Such artifacts typically occur around the poles and lead to a breakdown of the model (Figure 2).

You may now wonder, what would an FNO on a sphere look like?

GIF shows two spherical plots, comparing polar artifacts of AFNO to SFNO. Artifacts are not present with SFNO, while AFNO shows artifacts rapidly accumulating.
Figure 2. Temperature predictions using AFNO vs. SFNO

Figure 2 shows temperature predictions using adaptive Fourier neural operators (AFNO) as compared to spherical Fourier neural operators (SFNO). Respecting the spherical geometry and associated symmetries avoids artifacts and enables a stable rollout.

Spherical Fourier neural operators

To respect the spherical geometry of Earth, we implemented spherical Fourier neural operators (SFNOs), a Fourier neural operator that is directly formulated in spherical coordinates. To achieve this, we made use of a convolution theorem formulated on the sphere.

Global convolutions are the central building blocks of FNOs. Their computation through FFTs is enabled by the convolution theorem: a powerful mathematical tool that connects convolutions to the Fourier transform.

Similarly, a convolution theorem on the sphere connects spherical convolutions to the generalization of the Fourier transform on the sphere: the spherical harmonic transform (SHT).

Implementing differentiable spherical harmonic transforms

To enable the implementation of SFNOs, we required a differentiable SHT. To this end, we implemented torch-harmonics, a PyTorch library for differentiable SHTs. The library natively supports the computation of SHTs on single and multiple GPUs as well as CPUs, to enable scalable model parallelism. torch-harmonics can be installed easily by running the following command:

pip install torch-harmonics

torch-harmonics seamlessly integrates with PyTorch. The differentiable SHT can be easily integrated into any existing ML architecture as a module. To compute the SHT of a random function, run the following code example:

import torch
import torch_harmonics as th

device = torch.device('cuda' if torch.cuda.is_available() else 'cpu')

# parameters
nlat = 512
nlon = 2*nlat
batch_size = 32
signal = torch.randn(batch_size, nlat, nlon)

# create SHT instance
sht = th.RealSHT(nlat, nlon).to(device).float()

# execute transform
coeffs = sht(signal)

To get started with torch-harmonics, we recommend the getting-started notebook, which guides you through the computation of the spherical harmonic coefficients of Mars’ elevation map (Figure 3). The example showcases the computation of the coefficients using both the SHT and the differentiability of the ISHT.

Logarithmic chart depicting spherical harmonic coefficients next to an ellipsoid elevation map with color coding for elevation.
Figure 3. Spherical harmonic coefficients of the elevation map of Mars, computed with torch-harmonics (left). Reconstructed signal computed with the inverse spherical harmonic transform (right).

Implications for ML-based weather forecasting

We trained SFNOs on the ERA5 dataset, provided by the European Centre for Medium-range Weather Forecasts (ECMWF). This dataset represents our best understanding of the state of Earth’s atmosphere over the past 44 years. Figure 2 shows that SFNO shows no signs of artifacts over the poles and rollouts remain remarkably stable, over thousands of autoregressive steps, for up to a year (Figure 1).

These results pave the way for the deployment of ML-based weather prediction methods. They offer a glimpse of how ML-based methods may hold the key to bridging the gap between weather forecasting and climate prediction, in the holy grail of sub-seasonal-to-seasonal forecasting.

A single rollout of SFNOs for a year, which involves 1460 autoregressive steps, is computed in 13 minutes on a single NVIDIA RTX A6000. That is over a thousand times faster than traditional numerical weather prediction methods.

Such substantially faster forecasting tools open the door to the computation of thousands of possible scenarios in the same time that it took to do a single one using traditional NWP, enabling higher confidence predictions of the risk of rare but high-impact extreme weather events.

More about SFNOs and the NVIDIA Earth-2 initiative

To see how SFNOs were used to generate thousands of ensemble members and predict the 2018 Algerian heat wave, watch the following video:

Video 1. Predicting Extreme Weather Risk Three Weeks in Advance with FourCastNet

For more information about SFNOs, see the following resources:

For more information about the NVIDIA Earth-2 initiative, see the following resources:

Categories
Misc

How AI Is Powering the Future of Clean Energy

AI is improving ways to power the world by tapping the sun and the wind, along with cutting-edge technologies. The latest episode in the I AM AI video series showcases how artificial intelligence can help optimize solar and wind farms, simulate climate and weather, enhance power grid reliability and resilience, advance carbon capture and power Read article >

Categories
Misc

Gear Up and Game On: Gearbox’s ‘Remnant II’ Streaming on GeForce NOW

Get ready for Gunfire Games and Gearbox Publishing’s highly anticipated Remnant II, available for members to stream on GeForce NOW at launch. It leads eight new games coming to the cloud gaming platform. Ultimate and Priority members, make sure to grab the Guild Wars 2 rewards, available now through Thursday, Aug. 31. Visit the GeForce Read article >

Categories
Misc

ServiceNow, NVIDIA, and Accenture Team to Accelerate Generative AI Adoption for Enterprises

ServiceNow (NYSE: NOW), NVIDIA (NASDAQ: NVDA), and Accenture (NYSE: ACN) today announced the launch of AI Lighthouse, a first-of-its-kind program designed to fast-track the development and adoption of enterprise generative AI capabilities.

Categories
Misc

NVIDIA H100 GPUs Now Available on AWS Cloud

AWS users can now access the leading performance demonstrated in industry benchmarks of AI training and inference. The cloud giant officially switched on a new Amazon EC2 P5 instance powered by NVIDIA H100 Tensor Core GPUs. The service lets users scale generative AI, high performance computing (HPC) and other applications with a click from a Read article >

Categories
Offsites

In search of a generalizable method for source-free domain adaptation

Deep learning has recently made tremendous progress in a wide range of problems and applications, but models often fail unpredictably when deployed in unseen domains or distributions. Source-free domain adaptation (SFDA) is an area of research that aims to design methods for adapting a pre-trained model (trained on a “source domain”) to a new “target domain”, using only unlabeled data from the latter.

Designing adaptation methods for deep models is an important area of research. While the increasing scale of models and training datasets has been a key ingredient to their success, a negative consequence of this trend is that training such models is increasingly computationally expensive, out of reach for certain practitioners and also harmful for the environment. One avenue to mitigate this issue is through designing techniques that can leverage and reuse already trained models for tackling new tasks or generalizing to new domains. Indeed, adapting models to new tasks is widely studied under the umbrella of transfer learning.

SFDA is a particularly practical area of this research because several real-world applications where adaptation is desired suffer from the unavailability of labeled examples from the target domain. In fact, SFDA is enjoying increasing attention [1, 2, 3, 4]. However, albeit motivated by ambitious goals, most SFDA research is grounded in a very narrow framework, considering simple distribution shifts in image classification tasks.

In a significant departure from that trend, we turn our attention to the field of bioacoustics, where naturally-occurring distribution shifts are ubiquitous, often characterized by insufficient target labeled data, and represent an obstacle for practitioners. Studying SFDA in this application can, therefore, not only inform the academic community about the generalizability of existing methods and identify open research directions, but can also directly benefit practitioners in the field and aid in addressing one of the biggest challenges of our century: biodiversity preservation.

In this post, we announce “In Search for a Generalizable Method for Source-Free Domain Adaptation”, appearing at ICML 2023. We show that state-of-the-art SFDA methods can underperform or even collapse when confronted with realistic distribution shifts in bioacoustics. Furthermore, existing methods perform differently relative to each other than observed in vision benchmarks, and surprisingly, sometimes perform worse than no adaptation at all. We also propose NOTELA, a new simple method that outperforms existing methods on these shifts while exhibiting strong performance on a range of vision datasets. Overall, we conclude that evaluating SFDA methods (only) on the commonly-used datasets and distribution shifts leaves us with a myopic view of their relative performance and generalizability. To live up to their promise, SFDA methods need to be tested on a wider range of distribution shifts, and we advocate for considering naturally-occurring ones that can benefit high-impact applications.

Distribution shifts in bioacoustics

Naturally-occurring distribution shifts are ubiquitous in bioacoustics. The largest labeled dataset for bird songs is Xeno-Canto (XC), a collection of user-contributed recordings of wild birds from across the world. Recordings in XC are “focal”: they target an individual captured in natural conditions, where the song of the identified bird is at the foreground. For continuous monitoring and tracking purposes, though, practitioners are often more interested in identifying birds in passive recordings (“soundscapes”), obtained through omnidirectional microphones. This is a well-documented problem that recent work shows is very challenging. Inspired by this realistic application, we study SFDA in bioacoustics using a bird species classifier that was pre-trained on XC as the source model, and several “soundscapes” coming from different geographical locations — Sierra Nevada (S. Nevada); Powdermill Nature Reserve, Pennsylvania, USA; Hawai’i; Caples Watershed, California, USA; Sapsucker Woods, New York, USA (SSW); and Colombia — as our target domains.

This shift from the focalized to the passive domain is substantial: the recordings in the latter often feature much lower signal-to-noise ratio, several birds vocalizing at once, and significant distractors and environmental noise, like rain or wind. In addition, different soundscapes originate from different geographical locations, inducing extreme label shifts since a very small portion of the species in XC will appear in a given location. Moreover, as is common in real-world data, both the source and target domains are significantly class imbalanced, because some species are significantly more common than others. In addition, we consider a multi-label classification problem since there may be several birds identified within each recording, a significant departure from the standard single-label image classification scenario where SFDA is typically studied.

Illustration of the “focal → soundscapes” shift. In the focalized domain, recordings are typically composed of a single bird vocalization in the foreground, captured with high signal-to-noise ratio (SNR), though there may be other birds vocalizing in the background. On the other hand, soundscapes contain recordings from omnidirectional microphones and can be composed of multiple birds vocalizing simultaneously, as well as environmental noises from insects, rain, cars, planes, etc.

Audio files           

     Focal domain
     

     

     Soundscape domain1
     

Spectogram images                 
Illustration of the distribution shift from the focal domain (left) to the soundscape domain (right), in terms of the audio files (top) and spectrogram images (bottom) of a representative recording from each dataset. Note that in the second audio clip, the bird song is very faint; a common property in soundscape recordings where bird calls aren’t at the “foreground”. Credits: Left: XC recording by Sue Riffe (CC-BY-NC license). Right: Excerpt from a recording made available by Kahl, Charif, & Klinck. (2022) “A collection of fully-annotated soundscape recordings from the Northeastern United States” [link] from the SSW soundscape dataset (CC-BY license).

State-of-the-art SFDA models perform poorly on bioacoustics shifts

As a starting point, we benchmark six state-of-the-art SFDA methods on our bioacoustics benchmark, and compare them to the non-adapted baseline (the source model). Our findings are surprising: without exception, existing methods are unable to consistently outperform the source model on all target domains. In fact, they often underperform it significantly.

As an example, Tent, a recent method, aims to make models produce confident predictions for each example by reducing the uncertainty of the model’s output probabilities. While Tent performs well in various tasks, it doesn’t work effectively for our bioacoustics task. In the single-label scenario, minimizing entropy forces the model to choose a single class for each example confidently. However, in our multi-label scenario, there’s no such constraint that any class should be selected as being present. Combined with significant distribution shifts, this can cause the model to collapse, leading to zero probabilities for all classes. Other benchmarked methods like SHOT, AdaBN, Tent, NRC, DUST and Pseudo-Labelling, which are strong baselines for standard SFDA benchmarks, also struggle with this bioacoustics task.

Evolution of the test mean average precision (mAP), a standard metric for multilabel classification, throughout the adaptation procedure on the six soundscape datasets. We benchmark our proposed NOTELA and Dropout Student (see below), as well as SHOT, AdaBN, Tent, NRC, DUST and Pseudo-Labelling. Aside from NOTELA, all other methods fail to consistently improve the source model.

Introducing NOisy student TEacher with Laplacian Adjustment (NOTELA)

Nonetheless, a surprisingly positive result stands out: the less celebrated Noisy Student principle appears promising. This unsupervised approach encourages the model to reconstruct its own predictions on some target dataset, but under the application of random noise. While noise may be introduced through various channels, we strive for simplicity and use model dropout as the only noise source: we therefore refer to this approach as Dropout Student (DS). In a nutshell, it encourages the model to limit the influence of individual neurons (or filters) when making predictions on a specific target dataset.

DS, while effective, faces a model collapse issue on various target domains. We hypothesize this happens because the source model initially lacks confidence in those target domains. We propose improving DS stability by using the feature space directly as an auxiliary source of truth. NOTELA does this by encouraging similar pseudo-labels for nearby points in the feature space, inspired by NRC’s method and Laplacian regularization. This simple approach is visualized below, and consistently and significantly outperforms the source model in both audio and visual tasks.

NOTELA in action. The audio recordings are forwarded through the full model to obtain a first set of predictions, which are then refined through Laplacian regularization, a form of post-processing based on clustering nearby points. Finally, the refined predictions are used as targets for the noisy model to reconstruct.

Conclusion

The standard artificial image classification benchmarks have inadvertently limited our understanding of the true generalizability and robustness of SFDA methods. We advocate for broadening the scope and adopt a new assessment framework that incorporates naturally-occurring distribution shifts from bioacoustics. We also hope that NOTELA serves as a robust baseline to facilitate research in that direction. NOTELA’s strong performance perhaps points to two factors that can lead to developing more generalizable models: first, developing methods with an eye towards harder problems and second, favoring simple modeling principles. However, there is still future work to be done to pinpoint and comprehend existing methods’ failure modes on harder problems. We believe that our research represents a significant step in this direction, serving as a foundation for designing SFDA methods with greater generalizability.

Acknowledgements

One of the authors of this post, Eleni Triantafillou, is now at Google DeepMind. We are posting this blog post on behalf of the authors of the NOTELA paper: Malik Boudiaf, Tom Denton, Bart van Merriënboer, Vincent Dumoulin*, Eleni Triantafillou* (where * denotes equal contribution). We thank our co-authors for the hard work on this paper and the rest of the Perch team for their support and feedback.


1Note that in this audio clip, the bird song is very faint; a common property in soundscape recordings where bird calls aren’t at the “foreground”. 

Categories
Misc

Codeium’s Varun Mohan and Jeff Wang on Unleashing the Power of AI in Software Development

The world increasingly runs on code. Accelerating the work of those who create that code will boost their productivity — and that’s just what AI startup Codeium, a member of NVIDIA’s Inception program for startups, aims to do. On the latest episode of NVIDIA’s AI Podcast, host Noah Kravitz interviewed Codeium founder and CEO Varun Read article >

Categories
Misc

Ask Me Anything: NVIDIA CUDA Toolkit 12

CUDA Toolkit AMA promo card.On July 26, connect with CUDA product team experts on the latest NVIDIA CUDA Toolkit 12. CUDA Toolkit AMA promo card.

On July 26, connect with CUDA product team experts on the latest NVIDIA CUDA Toolkit 12.