Tag Archives: speech

Unsupervised speech-to-speech translation from monolingual data

Speech-to-speech translation (S2ST) is a type of machine translation that converts spoken language from one language to another. This technology has the potential to break down language barriers and facilitate communication between people from different cultures and backgrounds.

Previously, we introduced Translatotron 1 and Translatotron 2, the first ever models that were able to directly translate speech between two languages. However they were trained in supervised settings with parallel speech data. The scarcity of parallel speech data is a major challenge in this field, so much that most public datasets are semi- or fully-synthesized from text. This adds additional hurdles to learning translation and reconstruction of speech attributes that are not represented in the text and are thus not reflected in the synthesized training data.

Here we present Translatotron 3, a novel unsupervised speech-to-speech translation architecture. In Translatotron 3, we show that it is possible to learn a speech-to-speech translation task from monolingual data alone. This method opens the door not only to translation between more language pairs but also towards translation of the non-textual speech attributes such as pauses, speaking rates, and speaker identity. Our method does not include any direct supervision to target languages and therefore we believe it is the right direction for paralinguistic characteristics (e.g., such as tone, emotion) of the source speech to be preserved across translation. To enable speech-to-speech translation, we use back-translation, which is a technique from unsupervised machine translation (UMT) where a synthetic translation of the source language is used to translate texts without bilingual text datasets. Experimental results in speech-to-speech translation tasks between Spanish and English show that Translatotron 3 outperforms a baseline cascade system.


Translatotron 3

Translatotron 3 addresses the problem of unsupervised S2ST, which can eliminate the requirement for bilingual speech datasets. To do this, Translatotron 3’s design incorporates three key aspects:

  1. Pre-training the entire model as a masked autoencoder with SpecAugment, a simple data augmentation method for speech recognition that operates on the logarithmic mel spectogram of the input audio (instead of the raw audio itself) and is shown to effectively improve the generalization capabilities of the encoder.
  2. Unsupervised embedding mapping based on multilingual unsupervised embeddings (MUSE), which is trained on unpaired languages but allows the model to learn an embedding space that is shared between the source and target languages.
  3. A reconstruction loss based on back-translation, to train an encoder-decoder direct S2ST model in a fully unsupervised manner.

The model is trained using a combination of the unsupervised MUSE embedding loss, reconstruction loss, and S2S back-translation loss. During inference, the shared encoder is utilized to encode the input into a multilingual embedding space, which is subsequently decoded by the target language decoder.


Architecture

Translatotron 3 employs a shared encoder to encode both the source and target languages. The decoder is composed of a linguistic decoder, an acoustic synthesizer (responsible for acoustic generation of the translation speech), and a singular attention module, like Translatotron 2. However, for Translatotron 3 there are two decoders, one for the source language and another for the target language. During training, we use monolingual speech-text datasets (i.e., these data are made up of speech-text pairs; they are not translations).


Encoder

The encoder has the same architecture as the speech encoder in the Translatotron 2. The output of the encoder is split into two parts: the first part incorporates semantic information whereas the second part incorporates acoustic information. By using the MUSE loss, the first half of the output is trained to be the MUSE embeddings of the text of the input speech spectrogram. The latter half is updated without the MUSE loss. It is important to note that the same encoder is shared between source and target languages. Furthermore, the MUSE embedding is multilingual in nature. As a result, the encoder is able to learn a multilingual embedding space across source and target languages. This allows a more efficient and effective encoding of the input, as the encoder is able to encode speech from both languages into a common embedding space, rather than maintaining a separate embedding space for each language.


Decoder

Like Translatotron 2, the decoder is composed of three distinct components, namely the linguistic decoder, the acoustic synthesizer, and the attention module. To effectively handle the different properties of the source and target languages, however, Translatotron 3 has two separate decoders, for the source and target languages.


Two part training

The training methodology consists of two parts: (1) auto-encoding with reconstruction and (2) a back-translation term. In the first part, the network is trained to auto-encode the input to a multilingual embedding space using the MUSE loss and the reconstruction loss. This phase aims to ensure that the network generates meaningful multilingual representations. In the second part, the network is further trained to translate the input spectrogram by utilizing the back-translation loss. To mitigate the issue of catastrophic forgetting and enforcing the latent space to be multilingual, the MUSE loss and the reconstruction loss are also applied in this second part of training. To ensure that the encoder learns meaningful properties of the input, rather than simply reconstructing the input, we apply SpecAugment to encoder input at both phases. It has been shown to effectively improve the generalization capabilities of the encoder by augmenting the input data.


Training objective

During the back-translation training phase (illustrated in the section below), the network is trained to translate the input spectrogram to the target language and then back to the source language. The goal of back-translation is to enforce the latent space to be multilingual. To achieve this, the following losses are applied:

  • MUSE loss: The MUSE loss measures the similarity between the multilingual embedding of the input spectrogram and the multilingual embedding of the back-translated spectrogram.
  • Reconstruction loss: The reconstruction loss measures the similarity between the input spectrogram and the back-translated spectrogram.

In addition to these losses, SpecAugment is applied to the encoder input at both phases. Before the back-translation training phase, the network is trained to auto-encode the input to a multilingual embedding space using the MUSE loss and reconstruction loss.


MUSE loss

To ensure that the encoder generates multilingual representations that are meaningful for both decoders, we employ a MUSE loss during training. The MUSE loss forces the encoder to generate such a representation by using pre-trained MUSE embeddings. During the training process, given an input text transcript, we extract the corresponding MUSE embeddings from the embeddings of the input language. The error between MUSE embeddings and the output vectors of the encoder is then minimized. Note that the encoder is indifferent to the language of the input during inference due to the multilingual nature of the embeddings.

The training and inference in Translatotron 3. Training includes the reconstruction loss via the auto-encoding path and employs the reconstruction loss via back-translation.

Audio samples

Following are examples of direct speech-to-speech translation from Translatotron 3:

Spanish-to-English (on Conversational dataset)


Input (Spanish)
TTS-synthesized reference (English)   
Translatotron 3 (English)

Spanish-to-English (on CommonVoice11 Synthesized dataset)


Input (Spanish)
TTS-synthesized reference (English)   
Translatotron 3 (English)

Spanish-to-English (on CommonVoice11 dataset)


Input (Spanish)
TTS reference (English)
Translatotron 3 (English)   

Performance

To empirically evaluate the performance of the proposed approach, we conducted experiments on English and Spanish using various datasets, including the Common Voice 11 dataset, as well as two synthesized datasets derived from the Conversational and Common Voice 11 datasets.

The translation quality was measured by BLEU (higher is better) on ASR (automatic speech recognition) transcriptions from the translated speech, compared to the corresponding reference translation text. Whereas, the speech quality is measured by the MOS score (higher is better). Furthermore, the speaker similarity is measured by the average cosine similarity (higher is better).

Because Translatotron 3 is an unsupervised method, as a baseline we used a cascaded S2ST system that is combined from ASR, unsupervised machine translation (UMT), and TTS (text-to-speech). Specifically, we employ UMT that uses the nearest neighbor in the embedding space in order to create the translation.

Translatotron 3 outperforms the baseline by large margins in every aspect we measured: translation quality, speaker similarity, and speech quality. It particularly excelled on the conversational corpus. Moreover, Translatotron 3 achieves speech naturalness similar to that of the ground truth audio samples (measured by MOS, higher is better).

Translation quality (measured by BLEU, where higher is better) evaluated on three Spanish-English corpora.
Speech similarity (measured by average cosine similarity between input speaker and output speaker, where higher is better) evaluated on three Spanish-English corpora.
Mean-opinion-score (measured by average MOS metric, where higher is better) evaluated on three Spanish-English corpora.

Future work

As future work, we would like to extend the work to more languages and investigate whether zero-shot S2ST can be applied with the back-translation technique. We would also like to examine the use of back-translation with different types of speech data, such as noisy speech and low-resource languages.


Acknowledgments

The direct contributors to this work include Eliya Nachmani, Alon Levkovitch, Yifan Ding, Chulayutsh Asawaroengchai, Heiga Zhen, and Michelle Tadmor Ramanovich. We also thank Yu Zhang, Yuma Koizumi, Soroosh Mariooryad, RJ Skerry-Ryan, Neil Zeghidour, Christian Frank, Marco Tagliasacchi, Nadav Bar, Benny Schlesinger and Yonghui Wu.

Source: Google AI Blog


Spoken question answering and speech continuation using a spectrogram-powered LLM

The goal of natural language processing (NLP) is to develop computational models that can understand and generate natural language. By capturing the statistical patterns and structures of text-based natural language, language models can predict and generate coherent and meaningful sequences of words. Enabled by the increasing use of the highly successful Transformer model architecture and with training on large amounts of text (with proportionate compute and model size), large language models (LLMs) have demonstrated remarkable success in NLP tasks.

However, modeling spoken human language remains a challenging frontier. Spoken dialog systems have conventionally been built as a cascade of automatic speech recognition (ASR), natural language understanding (NLU), response generation, and text-to-speech (TTS) systems. However, to date there have been few capable end-to-end systems for the modeling of spoken language: i.e., single models that can take speech inputs and generate its continuation as speech outputs.

Today we present a new approach for spoken language modeling, called Spectron, published in “Spoken Question Answering and Speech Continuation Using Spectrogram-Powered LLM.” Spectron is the first spoken language model that is trained end-to-end to directly process spectrograms as both input and output, instead of learning discrete speech representations. Using only a pre-trained text language model, it can be fine-tuned to generate high-quality, semantically accurate spoken language. Furthermore, the proposed model improves upon direct initialization in retaining the knowledge of the original LLM as demonstrated through spoken question answering datasets.

We show that a pre-trained speech encoder and a language model decoder enable end-to-end training and state-of-the-art performance without sacrificing representational fidelity. Key to this is a novel end-to-end training objective that implicitly supervises speech recognition, text continuation, and conditional speech synthesis in a joint manner. A new spectrogram regression loss also supervises the model to match the higher-order derivatives of the spectrogram in the time and frequency domain. These derivatives express information aggregated from multiple frames at once. Thus, they express rich, longer-range information about the shape of the signal. Our overall scheme is summarized in the following figure:

The Spectron model connects the encoder of a speech recognition model with a pre-trained Transformer-based decoder language model. At training, speech utterances split into a prompt and its continuation. Then the full transcript (prompt and continuation) is reconstructed along with the continuation’s speech features. At inference, only a prompt is provided; the prompt’s transcription, text continuation, and speech continuations are all generated by the model.

Spectron architecture

The architecture is initialized with a pre-trained speech encoder and a pre-trained decoder language model. The encoder is prompted with a speech utterance as input, which it encodes into continuous linguistic features. These features feed into the decoder as a prefix, and the whole encoder-decoder is optimized to jointly minimize a cross-entropy loss (for speech recognition and transcript continuation) and a novel reconstruction loss (for speech continuation). During inference, one provides a spoken speech prompt, which is encoded and then decoded to give both text and speech continuations.


Speech encoder

The speech encoder is a 600M-parameter conformer encoder pre-trained on large-scale data (12M hours). It takes the spectrogram of the source speech as input, generating a hidden representation that incorporates both linguistic and acoustic information. The input spectrogram is first subsampled using a convolutional layer and then processed by a series of conformer blocks. Each conformer block consists of a feed-forward layer, a self-attention layer, a convolution layer, and a second feed-forward layer. The outputs are passed through a projection layer to match the hidden representations to the embedding dimension of the language model.


Language model

We use a 350M or 1B parameter decoder language model (for the continuation and question-answering tasks, respectively) trained in the manner of PaLM 2. The model receives the encoded features of the prompt as a prefix. Note that this is the only connection between the speech encoder and the LM decoder; i.e., there is no cross-attention between the encoder and the decoder. Unlike most spoken language models, during training, the decoder is teacher-forced to predict the text transcription, text continuation, and speech embeddings. To convert the speech embeddings to and from spectrograms, we introduce lightweight modules pre- and post-network.

By having the same architecture decode the intermediate text and the spectrograms, we gain two benefits. First, the pre-training of the LM in the text domain allows continuation of the prompt in the text domain before synthesizing the speech. Secondly, the predicted text serves as intermediate reasoning, enhancing the quality of the synthesized speech, analogous to improvements in text-based language models when using intermediate scratchpads or chain-of-thought (CoT) reasoning.


Acoustic projection layers

To enable the language model decoder to model spectrogram frames, we employ a multi-layer perceptron “pre-net” to project the ground truth spectrogram speech continuations to the language model dimension. This pre-net compresses the spectrogram input into a lower dimension, creating a bottleneck that aids the decoding process. This bottleneck mechanism prevents the model from repetitively generating the same prediction in the decoding process. To project the LM output from the language model dimension to the spectrogram dimension, the model employs a “post-net”, which is also a multi-layer perceptron. Both pre- and post-networks are two-layer multi-layer perceptrons.


Training objective

The training methodology of Spectron uses two distinct loss functions: (i) cross-entropy loss, employed for both speech recognition and transcript continuation, and (ii) regression loss, employed for speech continuation. During training, all parameters are updated (speech encoder, projection layer, LM, pre-net, and post-net).


Audio samples

Following are examples of speech continuation and question answering from Spectron:

Speech Continuation

Prompt:   
Continuation:   
  
Prompt:   
Continuation:   
  
Prompt:   
Continuation:   
  
Prompt:   
Continuation:   
  

Question Answering

Question:   
Answer:   
  
Question:   
Answer:   

Performance

To empirically evaluate the performance of the proposed approach, we conducted experiments on the Libri-Light dataset. Libri-Light is a 60k hour English dataset consisting of unlabelled speech readings from LibriVox audiobooks. We utilized a frozen neural vocoder called WaveFit to convert the predicted spectrograms into raw audio. We experiment with two tasks, speech continuation and spoken question answering (QA). Speech continuation quality is tested on the LibriSpeech test set. Spoken QA is tested on the Spoken WebQuestions datasets and a new test set named LLama questions, which we created. For all experiments, we use a 3 second audio prompt as input. We compare our method against existing spoken language models: AudioLM, GSLM, TWIST and SpeechGPT. For the speech continuation task, we use the 350M parameter version of LM and the 1B version for the spoken QA task.

For the speech continuation task, we evaluate our method using three metrics. The first is log-perplexity, which uses an LM to evaluate the cohesion and semantic quality of the generated speech. The second is mean opinion score (MOS), which measures how natural the speech sounds to human evaluators. The third, speaker similarity, uses a speaker encoder to measure how similar the speaker in the output is to the speaker in the input. Performance in all 3 metrics can be seen in the following graphs.

Log-perplexity for completions of LibriSpeech utterances given a 3-second prompt. Lower is better.
Speaker similarity between the prompt speech and the generated speech using the speaker encoder. Higher is better.
MOS given by human users on speech naturalness. Raters rate 5-scale subjective mean opinion score (MOS) ranging between 0 - 5 in naturalness given a speech utterance. Higher is better.

As can be seen in the first graph, our method significantly outperforms GSLM and TWIST on the log-perplexity metric, and does slightly better than state-of-the-art methods AudioLM and SpeechGPT. In terms of MOS, Spectron exceeds the performance of all the other methods except for AudioLM. In terms of speaker similarity, our method outperforms all other methods.

To evaluate the ability of the models to perform question answering, we use two spoken question answering datasets. The first is the LLama Questions dataset, which uses general knowledge questions in different domains generated using the LLama2 70B LLM. The second dataset is the WebQuestions dataset which is a general question answering dataset. For evaluation we use only questions that fit into the 3 second prompt length. To compute accuracy, answers are transcribed and compared to the ground truth answers in text form.

Accuracy for Question Answering on the LLama Questions and Spoken WebQuestions datasets. Accuracy is computed using the ASR transcripts of spoken answers.

First, we observe that all methods have more difficulty answering questions from the Spoken WebQuestions dataset than from the LLama questions dataset. Second, we observe that methods centered around spoken language modeling such as GSLM, AudioLM and TWIST have a completion-centric behavior rather than direct question answering which hindered their ability to perform QA. On the LLama questions dataset our method outperforms all other methods, while SpeechGPT is very close in performance. On the Spoken WebQuestions dataset, our method outperforms all other methods except for SpeechGPT, which does marginally better.


Acknowledgements

The direct contributors to this work include Eliya Nachmani, Alon Levkovitch, Julian Salazar, Chulayutsh Asawaroengchai, Soroosh Mariooryad, RJ Skerry-Ryan and Michelle Tadmor Ramanovich. We also thank Heiga Zhen, Yifan Ding, Yu Zhang, Yuma Koizumi, Neil Zeghidour, Christian Frank, Marco Tagliasacchi, Nadav Bar, Benny Schlesinger and Blaise Aguera-Arcas.

Source: Google AI Blog


English learners can now practice speaking on Search

Learning a language can open up new opportunities in a person’s life. It can help people connect with those from different cultures, travel the world, and advance their career. English alone is estimated to have 1.5 billion learners worldwide. Yet proficiency in a new language is difficult to achieve, and many learners cite a lack of opportunity to practice speaking actively and receiving actionable feedback as a barrier to learning.

We are excited to announce a new feature of Google Search that helps people practice speaking and improve their language skills. Within the next few days, Android users in Argentina, Colombia, India (Hindi), Indonesia, Mexico, and Venezuela can get even more language support from Google through interactive speaking practice in English — expanding to more countries and languages in the future. Google Search is already a valuable tool for language learners, providing translations, definitions, and other resources to improve vocabulary. Now, learners translating to or from English on their Android phones will find a new English speaking practice experience with personalized feedback.

A new feature of Google Search allows learners
to practice speaking words in context.

Learners are presented with real-life prompts and then form their own spoken answers using a provided vocabulary word. They engage in practice sessions of 3-5 minutes, getting personalized feedback and the option to sign up for daily reminders to keep practicing. With only a smartphone and some quality time, learners can practice at their own pace, anytime, anywhere.


Activities with personalized feedback, to supplement existing learning tools

Designed to be used alongside other learning services and resources, like personal tutoring, mobile apps, and classes, the new speaking practice feature on Google Search is another tool to assist learners on their journey.

We have partnered with linguists, teachers, and ESL/EFL pedagogical experts to create a speaking practice experience that is effective and motivating. Learners practice vocabulary in authentic contexts, and material is repeated over dynamic intervals to increase retention — approaches that are known to be effective in helping learners become confident speakers. As one partner of ours shared:

"Speaking in a given context is a skill that language learners often lack the opportunity to practice. Therefore this tool is very useful to complement classes and other resources." - Judit Kormos, Professor, Lancaster University

We are also excited to be working with several language learning partners to surface content they are helping create and to connect them with learners around the world. We look forward to expanding this program further and working with any interested partner.


Personalized real-time feedback

Every learner is different, so delivering personalized feedback in real time is a key part of effective practice. Responses are analyzed to provide helpful, real-time suggestions and corrections.

The system gives semantic feedback, indicating whether their response was relevant to the question and may be understood by a conversation partner. Grammar feedback provides insights into possible grammatical improvements, and a set of example answers at varying levels of language complexity give concrete suggestions for alternative ways to respond in this context.

The feedback is composed of three elements: Semantic analysis, grammar correction, and example answers.

Contextual translation

Among the several new technologies we developed, contextual translation provides the ability to translate individual words and phrases in context. During practice sessions, learners can tap on any word they don’t understand to see the translation of that word considering its context.

Example of contextual translation feature.

This is a difficult technical task, since individual words in isolation often have multiple alternative meanings, and multiple words can form clusters of meaning that need to be translated in unison. Our novel approach translates the entire sentence, then estimates how the words in the original and the translated text relate to each other. This is commonly known as the word alignment problem.

Example of a translated sentence pair and its word alignment. A deep learning alignment model connects the different words that create the meaning to suggest a translation.

The key technology piece that enables this functionality is a novel deep learning model developed in collaboration with the Google Translate team, called Deep Aligner. The basic idea is to take a multilingual language model trained on hundreds of languages, then fine-tune a novel alignment model on a set of word alignment examples (see the figure above for an example) provided by human experts, for several language pairs. From this, the single model can then accurately align any language pair, reaching state-of-the-art alignment error rate (AER, a metric to measure the quality of word alignments, where lower is better). This single new model has led to dramatic improvements in alignment quality across all tested language pairs, reducing average AER from 25% to 5% compared to alignment approaches based on Hidden Markov models (HMMs).

Alignment error rates (lower is better) between English (EN) and other languages.

This model is also incorporated into Google’s translation APIs, greatly improving, for example, the formatting of translated PDFs and websites in Chrome, the translation of YouTube captions, and enhancing Google Cloud’s translation API.


Grammar feedback

To enable grammar feedback for accented spoken language, our research teams adapted grammar correction models for written text (see the blog and paper) to work on automatic speech recognition (ASR) transcriptions, specifically for the case of accented speech. The key step was fine-tuning the written text model on a corpus of human and ASR transcripts of accented speech, with expert-provided grammar corrections. Furthermore, inspired by previous work, the teams developed a novel edit-based output representation that leverages the high overlap between the inputs and outputs that is particularly well-suited for short input sentences common in language learning settings.

The edit representation can be explained using an example:

  • Input: I1 am2 so3 bad4 cooking5
  • Correction: I1 am2 so3 bad4 at5 cooking6
  • Edits: ('at', 4, PREPOSITION, 4)

In the above, “at” is the word that is inserted at position 4 and “PREPOSITION” denotes this is an error involving prepositions. We used the error tag to select tag-dependent acceptance thresholds that improved the model further. The model increased the recall of grammar problems from 4.6% to 35%.

Some example output from our model and a model trained on written corpora:


    Example 1     Example 2
User input (transcribed speech) I live of my profession. I need a efficient card and reliable.
Text-based grammar model I live by my profession. I need an efficient card and a reliable.
New speech-optimized model I live off my profession. I need an efficient and reliable card.


Semantic analysis

A primary goal of conversation is to communicate one’s intent clearly. Thus, we designed a feature that visually communicates to the learner whether their response was relevant to the context and would be understood by a partner. This is a difficult technical problem, since early language learners’ spoken responses can be syntactically unconventional. We had to carefully balance this technology to focus on the clarity of intent rather than correctness of syntax.

Our system utilizes a combination of two approaches:

  1. Sensibility classification: Large language models like LaMDA or PaLM are designed to give natural responses in a conversation, so it’s no surprise that they do well on the reverse: judging whether a given response is contextually sensible.
  2. Similarity to good responses: We used an encoder architecture to compare the learner’s input to a set of known good responses in a semantic embedding space. This comparison provides another useful signal on semantic relevance, further improving the quality of feedback and suggestions we provide.
The system provides feedback about whether the response was relevant to the prompt, and would be understood by a communication partner.

ML-assisted content development

Our available practice activities present a mix of human-expert created content, and content that was created with AI assistance and human review. This includes speaking prompts, focus words, as well as sets of example answers that showcase meaningful and contextual responses.

A list of example answers is provided when the learner receives feedback and when they tap the help button.

Since learners have different levels of ability, the language complexity of the content has to be adjusted appropriately. Prior work on language complexity estimation focuses on text of paragraph length or longer, which differs significantly from the type of responses that our system processes. Thus, we developed novel models that can estimate the complexity of a single sentence, phrase, or even individual words. This is challenging because even a phrase composed of simple words can be hard for a language learner (e.g., "Let's cut to the chase”). Our best model is based on BERT and achieves complexity predictions closest to human expert consensus. The model was pre-trained using a large set of LLM-labeled examples, and then fine-tuned using a human expert–labeled dataset.

Mean squared error of various approaches’ performance estimating content difficulty on a diverse corpus of ~450 conversational passages (text / transcriptions). Top row: Human raters labeled the items on a scale from 0.0 to 5.0, roughly aligned to the CEFR scale (from A1 to C2). Bottom four rows: Different models performed the same task, and we show the difference to the human expert consensus.

Using this model, we can evaluate the difficulty of text items, offer a diverse range of suggestions, and most importantly challenge learners appropriately for their ability levels. For example, using our model to label examples, we can fine-tune our system to generate speaking prompts at various language complexity levels.


Vocabulary focus words, to be elicited by the questions
    guitar     apple     lion
Simple     What do you like to play?     Do you like fruit?     Do you like big cats?
Intermediate     Do you play any musical instruments?     What is your favorite fruit?     What is your favorite animal?
Complex     What stringed instrument do you enjoy playing?     Which type of fruit do you enjoy eating for its crunchy texture and sweet flavor?     Do you enjoy watching large, powerful predators?

Furthermore, content difficulty estimation is used to gradually increase the task difficulty over time, adapting to the learner’s progress.


Conclusion

With these latest updates, which will roll out over the next few days, Google Search has become even more helpful. If you are an Android user in India (Hindi), Indonesia, Argentina, Colombia, Mexico, or Venezuela, give it a try by translating to or from English with Google.

We look forward to expanding to more countries and languages in the future, and to start offering partner practice content soon.


Acknowledgements

Many people were involved in the development of this project. Among many others, we thank our external advisers in the language learning field: Jeffrey Davitz, Judit Kormos, Deborah Healey, Anita Bowles, Susan Gaer, Andrea Revesz, Bradley Opatz, and Anne Mcquade.

Source: Google AI Blog


SoundStorm: Efficient parallel audio generation

The recent progress in generative AI unlocked the possibility of creating new content in several different domains, including text, vision and audio. These models often rely on the fact that raw data is first converted to a compressed format as a sequence of tokens. In the case of audio, neural audio codecs (e.g., SoundStream or EnCodec) can efficiently compress waveforms to a compact representation, which can be inverted to reconstruct an approximation of the original audio signal. Such a representation consists of a sequence of discrete audio tokens, capturing the local properties of sounds (e.g., phonemes) and their temporal structure (e.g., prosody). By representing audio as a sequence of discrete tokens, audio generation can be performed with Transformer-based sequence-to-sequence models — this has unlocked rapid progress in speech continuation (e.g., with AudioLM), text-to-speech (e.g., with SPEAR-TTS), and general audio and music generation (e.g., AudioGen and MusicLM). Many generative audio models, including AudioLM, rely on auto-regressive decoding, which produces tokens one by one. While this method achieves high acoustic quality, inference (i.e., calculating an output) can be slow, especially when decoding long sequences.

To address this issue, in “SoundStorm: Efficient Parallel Audio Generation”, we propose a new method for efficient and high-quality audio generation. SoundStorm addresses the problem of generating long audio token sequences by relying on two novel elements: 1) an architecture adapted to the specific nature of audio tokens as produced by the SoundStream neural codec, and 2) a decoding scheme inspired by MaskGIT, a recently proposed method for image generation, which is tailored to operate on audio tokens. Compared to the autoregressive decoding approach of AudioLM, SoundStorm is able to generate tokens in parallel, thus decreasing the inference time by 100x for long sequences, and produces audio of the same quality and with higher consistency in voice and acoustic conditions. Moreover, we show that SoundStorm, coupled with the text-to-semantic modeling stage of SPEAR-TTS, can synthesize high-quality, natural dialogues, allowing one to control the spoken content (via transcripts), speaker voices (via short voice prompts) and speaker turns (via transcript annotations), as demonstrated by the examples below:





Input: Text (transcript used to drive the audio generation in bold)        Something really funny happened to me this morning. | Oh wow, what? | Well, uh I woke up as usual. | Uhhuh | Went downstairs to have uh breakfast. | Yeah | Started eating. Then uh 10 minutes later I realized it was the middle of the night. | Oh no way, that's so funny!        I didn't sleep well last night. | Oh, no. What happened? | I don't know. I I just couldn't seem to uh to fall asleep somehow, I kept tossing and turning all night. | That's too bad. Maybe you should uh try going to bed earlier tonight or uh maybe you could try reading a book. | Yeah, thanks for the suggestions, I hope you're right. | No problem. I I hope you get a good night's sleep
         
Input: Audio prompt         
         
Output: Audio prompt + generated audio              


SoundStorm design

In our previous work on AudioLM, we showed that audio generation can be decomposed into two steps: 1) semantic modeling, which generates semantic tokens from either previous semantic tokens or a conditioning signal (e.g., a transcript as in SPEAR-TTS, or a text prompt as in MusicLM), and 2) acoustic modeling, which generates acoustic tokens from semantic tokens. With SoundStorm we specifically address this second, acoustic modeling step, replacing slower autoregressive decoding with faster parallel decoding.

SoundStorm relies on a bidirectional attention-based Conformer, a model architecture that combines a Transformer with convolutions to capture both local and global structure of a sequence of tokens. Specifically, the model is trained to predict audio tokens produced by SoundStream given a sequence of semantic tokens generated by AudioLM as input. When doing this, it is important to take into account the fact that, at each time step t, SoundStream uses up to Q tokens to represent the audio using a method known as residual vector quantization (RVQ), as illustrated below on the right. The key intuition is that the quality of the reconstructed audio progressively increases as the number of generated tokens at each step goes from 1 to Q.

At inference time, given the semantic tokens as input conditioning signal, SoundStorm starts with all audio tokens masked out, and fills in the masked tokens over multiple iterations, starting from the coarse tokens at RVQ level q = 1 and proceeding level-by-level with finer tokens until reaching level q = Q.

There are two crucial aspects of SoundStorm that enable fast generation: 1) tokens are predicted in parallel during a single iteration within a RVQ level and, 2) the model architecture is designed in such a way that the complexity is only mildly affected by the number of levels Q. To support this inference scheme, during training a carefully designed masking scheme is used to mimic the iterative process used at inference.

SoundStorm model architecture. T denotes the number of time steps and Q the number of RVQ levels used by SoundStream. The semantic tokens used as conditioning are time-aligned with the SoundStream frames.

Measuring SoundStorm performance

We demonstrate that SoundStorm matches the quality of AudioLM's acoustic generator, replacing both AudioLM's stage two (coarse acoustic model) and stage three (fine acoustic model). Furthermore, SoundStorm produces audio 100x faster than AudioLM's hierarchical autoregressive acoustic generator (top half below) with matching quality and improved consistency in terms of speaker identity and acoustic conditions (bottom half below).

Runtimes of SoundStream decoding, SoundStorm and different stages of AudioLM on a TPU-v4.
Acoustic consistency between the prompt and the generated audio. The shaded area represents the inter-quartile range.

Safety and risk mitigation

We acknowledge that the audio samples produced by the model may be influenced by the unfair biases present in the training data, for instance in terms of represented accents and voice characteristics. In our generated samples, we demonstrate that we can reliably and responsibly control speaker characteristics via prompting, with the goal of avoiding unfair biases. A thorough analysis of any training data and its limitations is an area of future work in line with our responsible AI Principles.

In turn, the ability to mimic a voice can have numerous malicious applications, including bypassing biometric identification and using the model for the purpose of impersonation. Thus, it is crucial to put in place safeguards against potential misuse: to this end, we have verified that the audio generated by SoundStorm remains detectable by a dedicated classifier using the same classifier as described in our original AudioLM paper. Hence, as a component of a larger system, we believe that SoundStorm would be unlikely to introduce additional risks to those discussed in our earlier papers on AudioLM and SPEAR-TTS. At the same time, relaxing the memory and computational requirements of AudioLM would make research in the domain of audio generation more accessible to a wider community. In the future, we plan to explore other approaches for detecting synthesized speech, e.g., with the help of audio watermarking, so that any potential product usage of this technology strictly follows our responsible AI Principles.


Conclusion

We have introduced SoundStorm, a model that can efficiently synthesize high-quality audio from discrete conditioning tokens. When compared to the acoustic generator of AudioLM, SoundStorm is two orders of magnitude faster and achieves higher temporal consistency when generating long audio samples. By combining a text-to-semantic token model similar to SPEAR-TTS with SoundStorm, we can scale text-to-speech synthesis to longer contexts and generate natural dialogues with multiple speaker turns, controlling both the voices of the speakers and the generated content. SoundStorm is not limited to generating speech. For example, MusicLM uses SoundStorm to synthesize longer outputs efficiently (as seen at I/O).


Acknowledgments

The work described here was authored by Zalán Borsos, Matt Sharifi, Damien Vincent, Eugene Kharitonov, Neil Zeghidour and Marco Tagliasacchi. We are grateful for all discussions and feedback on this work that we received from our colleagues at Google.

Source: Google AI Blog


Universal Speech Model (USM): State-of-the-art speech AI for 100+ languages

Last November, we announced the 1,000 Languages Initiative, an ambitious commitment to build a machine learning (ML) model that would support the world’s one thousand most-spoken languages, bringing greater inclusion to billions of people around the globe. However, some of these languages are spoken by fewer than twenty million people, so a core challenge is how to support languages for which there are relatively few speakers or limited available data.

Today, we are excited to share more about the Universal Speech Model (USM), a critical first step towards supporting 1,000 languages. USM is a family of state-of-the-art speech models with 2B parameters trained on 12 million hours of speech and 28 billion sentences of text, spanning 300+ languages. USM, which is for use in YouTube (e.g., for closed captions), can perform automatic speech recognition (ASR) not only on widely-spoken languages like English and Mandarin, but also on under-resourced languages like Amharic, Cebuano, Assamese, and Azerbaijani to name a few. In “Google USM: Scaling Automatic Speech Recognition Beyond 100 Languages”, we demonstrate that utilizing a large unlabeled multilingual dataset to pre-train the encoder of the model and fine-tuning on a smaller set of labeled data enables us to recognize under-represented languages. Moreover, our model training process is effective at adapting to new languages and data.

A sample of the languages that USM supports.

Challenges in current ASR

To accomplish this ambitious goal, we need to address two significant challenges in ASR.

First, there is a lack of scalability with conventional supervised learning approaches. A fundamental challenge of scaling speech technologies to many languages is obtaining enough data to train high-quality models. With conventional approaches, audio data needs to be either manually labeled, which is time-consuming and costly, or collected from sources with pre-existing transcriptions, which are harder to find for languages that lack wide representation. In contrast, self-supervised learning can leverage audio-only data, which is available in much larger quantities across languages. This makes self-supervision a better approach to accomplish our goal of scaling across hundreds of languages.

Another challenge is that models must improve in a computationally efficient manner while we expand the language coverage and quality. This requires the learning algorithm to be flexible, efficient, and generalizable. More specifically, such an algorithm should be able to use large amounts of data from a variety of sources, enable model updates without requiring complete retraining, and generalize to new languages and use cases.


Our approach: Self-supervised learning with fine-tuning

USM uses the standard encoder-decoder architecture, where the decoder can be CTC, RNN-T, or LAS. For the encoder, USM uses the Conformer, or convolution-augmented transformer. The key component of the Conformer is the Conformer block, which consists of attention, feed-forward, and convolutional modules. It takes as input the log-mel spectrogram of the speech signal and performs a convolutional sub-sampling, after which a series of Conformer blocks and a projection layer are applied to obtain the final embeddings.

Our training pipeline starts with the first step of self-supervised learning on speech audio covering hundreds of languages. In the second optional step, the model’s quality and language coverage can be improved through an additional pre-training step with text data. The decision to incorporate the second step depends on whether text data is available. USM performs best with this second optional step. The last step of the training pipeline is to fine-tune on downstream tasks (e.g., ASR or automatic speech translation) with a small amount of supervised data.

For the first step, we use BEST-RQ, which has already demonstrated state-of-the-art results on multilingual tasks and has proven to be efficient when using very large amounts of unsupervised audio data.

In the second (optional) step, we used multi-objective supervised pre-training to incorporate knowledge from additional text data. The model introduces an additional encoder module to take text as input and additional layers to combine the output of the speech encoder and the text encoder, and trains the model jointly on unlabeled speech, labeled speech, and text data.

In the last stage, USM is fine-tuned on the downstream tasks. The overall training pipeline is illustrated below. With the knowledge acquired during pre-training, USM models achieve good quality with only a small amount of supervised data from the downstream tasks.

USM’s overall training pipeline.

Key results

Performance across multiple languages on YouTube Captions

Our encoder incorporates 300+ languages through pre-training. We demonstrate the effectiveness of the pre-trained encoder through fine-tuning on YouTube Caption’s multilingual speech data. The supervised YouTube data includes 73 languages and has on average less than three thousand hours of data per language. Despite limited supervised data, the model achieves less than 30% word error rate (WER; lower is better) on average across the 73 languages, a milestone we have never achieved before. For en-US, USM has a 6% relative lower WER compared to the current internal state-of-the-art model. Lastly, we compare with the recently released large model, Whisper (large-v2), which was trained with more than 400k hours of labeled data. For the comparison, we only use the 18 languages that Whisper can successfully decode with lower than 40% WER. Our model has, on average, a 32.7% relative lower WER compared to Whisper for these 18 languages.

USM supports all 73 languages in the YouTube Captions' Test Set and outperforms Whisper on the languages it can support with lower than 40% WER. Lower WER is better.

Generalization to downstream ASR tasks

On publicly available datasets, our model shows lower WER compared to Whisper on CORAAL (African American Vernacular English), SpeechStew (en-US), and FLEURS (102 languages). Our model achieves lower WER with and without training on in-domain data. The comparison on FLEURS reports the subset of languages (62) that overlaps with the languages supported by the Whisper model. For FLEURS, USM without in-domain data has a 65.8% relative lower WER compared to Whisper and has a 67.8% relative lower WER with in-domain data.

Comparison of USM (with or without in-domain data) and Whisper results on ASR benchmarks. Lower WER is better.

Performance on automatic speech translation (AST)

For speech translation, we fine-tune USM on the CoVoST dataset. Our model, which includes text via the second stage of our pipeline, achieves state-of-the-art quality with limited supervised data. To assess the breadth of the model’s performance, we segment the languages from the CoVoST dataset into high, medium, and low based on resource availability and calculate the BLEU score (higher is better) for each segment. As shown below, USM outperforms Whisper for all segments.

CoVoST BLEU score. Higher BLEU is better.

Toward 1,000 languages

The development of USM is a critical effort towards realizing Google’s mission to organize the world’s information and make it universally accessible. We believe USM’s base model architecture and training pipeline comprise a foundation on which we can build to expand speech modeling to the next 1,000 languages.


Learn More

Check out our paper here. Researchers can request access to the USM API here.


Acknowledgements

We thank all the co-authors for contributing to the project and paper, including Andrew Rosenberg, Ankur Bapna, Bhuvana Ramabhadran, Bo Li, Chung-Cheng Chiu, Daniel Park, Françoise Beaufays, Hagen Soltau, Gary Wang, Ginger Perng, James Qin, Jason Riesa, Johan Schalkwyk, Ke Hu, Nanxin Chen, Parisa Haghani, Pedro Moreno Mengibar, Rohit Prabhavalkar, Tara Sainath, Trevor Strohman, Vera Axelrod, Wei Han, Yonghui Wu, Yongqiang Wang, Yu Zhang, Zhehuai Chen, and Zhong Meng.

We also thank Alexis Conneau, Min Ma, Shikhar Bharadwaj, Sid Dalmia, Jiahui Yu, Jian Cheng, Paul Rubenstein, Ye Jia, Justin Snyder, Vincent Tsang, Yuanzhong Xu, Tao Wang for useful discussions.

We appreciate valuable feedback and support from Eli Collins, Jeff Dean, Sissie Hsiao, Zoubin Ghahramani. Special thanks to Austin Tarango, Lara Tumeh, Amna Latif, and Jason Porta for their guidance around Responsible AI practices. We thank Elizabeth Adkison, James Cokerille for help with naming the model, Tom Small for the animated graphic, Abhishek Bapna for editorial support, and Erica Moreira for resource management . We thank Anusha Ramesh for feedback, guidance, and assistance with the publication strategy, and Calum Barnes and Salem Haykal for their valuable partnership.

Source: Google AI Blog


Who Said What? Recorder’s On-device Solution for Labeling Speakers

In 2019 we launched Recorder, an audio recording app for Pixel phones that helps users create, manage, and edit audio recordings. It leverages recent developments in on-device machine learning to transcribe speech, recognize audio events, suggest tags for titles, and help users navigate transcripts.

Nonetheless, some Recorder users found it difficult to navigate long recordings that have multiple speakers because it's not clear who said what. During the Made By Google event this year, we announced the "speaker labels" feature for the Recorder app. This opt-in feature annotates a recording transcript with unique and anonymous labels for each speaker (e.g., "Speaker 1", "Speaker 2", etc.) in real time during the recording. It significantly improves the readability and usability of the recording transcripts. This feature is powered by Google's new speaker diarization system named Turn-to-Diarize, which was first presented at ICASSP 2022.

Left: Recorder transcript without speaker labels. Right: Recorder transcript with speaker labels.

System Architecture

Our speaker diarization system leverages several highly optimized machine learning models and algorithms to allow diarizing hours of audio in a real-time streaming fashion with limited computational resources on mobile devices. The system mainly consists of three components: a speaker turn detection model that detects a change of speaker in the input speech, a speaker encoder model that extracts voice characteristics from each speaker turn, and a multi-stage clustering algorithm that annotates speaker labels to each speaker turn in a highly efficient way. All components run fully on the device.

Architecture of the Turn-to-Diarize system.

Detecting Speaker Turns

The first component of our system is a speaker turn detection model based on a Transformer Transducer (T-T), which converts the acoustic features into text transcripts augmented with a special token <st> representing a speaker turn. Unlike preceding customized systems that use role-specific tokens (e.g., <doctor> and <patient>) for conversations, this model is more generic and can be trained on and deployed to various application domains.

In most applications, the output of a diarization system is not directly shown to users, but combined with a separate automatic speech recognition (ASR) system that is trained to have smaller word errors. Therefore, for the diarization system, we are relatively more tolerant to word token errors than errors of the <st> token. Based on this intuition, we propose a new token-level loss function that allows us to train a small speaker turn detection model with high accuracy on predicted <st> tokens. Combined with edit-based minimum Bayes risk (EMBR) training, this new loss function significantly improved the interval-based F1 score on seven evaluation datasets.


Extracting Voice Characteristics

Once the audio recording has been segmented into homogeneous speaker turns, we use a speaker encoder model to extract an embedding vector (i.e., d-vector) to represent the voice characteristics of each speaker turn. This approach has several advantages over prior work that extracts embedding vectors from small fixed-length segments. First, it avoids extracting an embedding from a segment containing speech from multiple speakers. At the same time, each embedding covers a relatively large time range that contains sufficient signals from the speaker. It also reduces the total number of embeddings to be clustered, thus making the clustering step less expensive. These embeddings are processed entirely on-device until speaker labeling of the transcript is completed, and then deleted.


Multi-Stage Clustering

After the audio recording is represented by a sequence of embedding vectors, the last step is to cluster these embedding vectors, and assign a speaker label to each. However, since audio recordings from the Recorder app can be as short as a few seconds, or as long as up to 18 hours, it is critical for the clustering algorithm to handle sequences of drastically different lengths.

For this we propose a multi-stage clustering strategy to leverage the benefits of different clustering algorithms. First, we use the speaker turn detection outputs to determine whether there are at least two different speakers in the recording. For short sequences, we use agglomerative hierarchical clustering (AHC) as the fallback algorithm. For medium-length sequences, we use spectral clustering as our main algorithm, and use the eigen-gap criterion for accurate speaker count estimation. For long sequences, we reduce computational cost by using AHC to pre-cluster the sequence before feeding it to the main algorithm. During the streaming, we keep a dynamic cache of previous AHC cluster centroids that can be reused for future clustering calls. This mechanism allows us to enforce an upper bound on the entire system with constant time and space complexity.

This multi-stage clustering strategy is a critical optimization for on-device applications where the budget for CPU, memory, and battery is very small, and allows the system to run in a low power mode even after diarizing hours of audio. As a tradeoff between quality and efficiency, the upper bound of the computational cost can be flexibly configured for devices with different computational resources.

Diagram of the multi-stage clustering strategy.

Correction and Customization

In our real-time streaming speaker diarization system, as the model consumes more audio input, it accumulates confidence on predicted speaker labels, and may occasionally make corrections to previously predicted low-confidence speaker labels. The Recorder app automatically updates the speaker labels on the screen during recording to reflect the latest and most accurate predictions.

At the same time, the Recorder app's UI allows the user to rename the anonymous speaker labels (e.g., "Speaker 2") to customized labels (e.g., "car dealer") for better readability and easier memorization for the user within each recording.

Recorder allows the user to rename the speaker labels for better readability.

Future Work

Currently, our diarization system mostly runs on the CPU block of Google Tensor, Google's custom-built chip that powers more recent Pixel phones. We are working on delegating more computations to the TPU block, which will further reduce the overall power consumption of the diarization system. Another future work direction is to leverage multilingual capabilities of speaker encoder and speech recognition models to expand this feature to more languages.


Acknowledgments

The work described in this post represents joint efforts from multiple teams within Google. Contributors include Quan Wang, Yiling Huang, Evan Clark, Qi Cao, Han Lu, Guanlong Zhao, Wei Xia, Hasim Sak, Alvin Zhou, Jason Pelecanos, Luiza Timariu, Allen Su, Fan Zhang, Hugh Love, Kristi Bradford, Vincent Peng, Raff Tsai, Richard Chou, Yitong Lin, Ann Lu, Kelly Tsai, Hannah Bowman, Tracy Wu, Taral Joglekar, Dharmesh Mokani, Ajay Dudani, Ignacio Lopez Moreno, Diego Melendo Casado, Nino Tasca, Alex Gruenstein.

Source: Google AI Blog


Google at Interspeech 2022

This week, the 23rd Annual Conference of the International Speech Communication Association (INTERSPEECH 2022) is being held in Incheon, South Korea, representing one of the world’s most extensive conferences on research and technology of spoken language understanding and processing. Over 2,000 experts in speech-related research fields gather to take part in oral presentations and poster sessions and to collaborate with streamed events across the globe.

We are excited to be a Diamond Sponsor of INTERSPEECH 2022, where we will be showcasing nearly 50 research publications and supporting a number of workshops, special sessions and tutorials. We welcome in-person attendees to drop by the Google booth to meet our researchers and participate in Q&As and demonstrations of some of our latest speech technologies, which help to improve accessibility and provide convenience in communication for billions of users. In addition, online attendees are encouraged to visit our virtual booth in GatherTown where you can get up-to-date information on research and opportunities at Google. You can also learn more about the Google research being presented at INTERSPEECH 2022 below (Google affiliations in bold).


Organizing Committee

Industry Liaisons include: Bhuvana Ramabahdran

Area Chairs include: John Hershey, Heiga Zen, Shrikanth Narayanan, Bastiaan Kleijn


ISCA Fellows

Include: Tara Sainath, Heiga Zen


Publications

Production Federated Keyword Spotting via Distillation, Filtering, and Joint Federated-Centralized Training
Andrew Hard, Kurt Partridge, Neng Chen, Sean Augenstein, Aishanee Shah, Hyun Jin Park, Alex Park, Sara Ng, Jessica Nguyen, Ignacio Lopez Moreno, Rajiv Mathews, Françoise Beaufays

Leveraging Unsupervised and Weakly-Supervised Data to Improve Direct Speech-to-Speech Translation
Ye Jia, Yifan Ding, Ankur Bapna, Colin Cherry, Yu Zhang, Alexis Conneau, Nobu Morioka

Sentence-Select: Large-Scale Language Model Data Selection for Rare-Word Speech Recognition
W. Ronny Huang, Cal Peyser, Tara N. Sainath, Ruoming Pang, Trevor Strohman, Shankar Kumar

UserLibri: A Dataset for ASR Personalization Using Only Text
Theresa Breiner, Swaroop Ramaswamy, Ehsan Variani, Shefali Garg, Rajiv Mathews, Khe Chai Sim, Kilol Gupta, Mingqing Chen, Lara McConnaughey

SNRi Target Training for Joint Speech Enhancement and Recognition
Yuma Koizumi, Shigeki Karita, Arun Narayanan, Sankaran Panchapagesan, Michiel Bacchiani

Turn-Taking Prediction for Natural Conversational Speech
Shuo-Yiin Chang, Bo Li, Tara Sainath, Chao Zhang, Trevor Strohman, Qiao Liang, Yanzhang He

Streaming Intended Query Detection Using E2E Modeling for Continued Conversation
Shuo-Yiin Chang, Guru Prakash, Zelin Wu, Tara Sainath, Bo Li, Qiao Liang, Adam Stambler, Shyam Upadhyay, Manaal Faruqui, Trevor Strohman

Improving Distortion Robustness of Self-Supervised Speech Processing Tasks with Domain Adaptation
Kuan Po Huang, Yu-Kuan Fu, Yu Zhang, Hung-yi Lee

XLS-R: Self-Supervised Cross-Lingual Speech Representation Learning at Scale
Arun Babu, Changhan Wang, Andros Tjandra, Kushal Lakhotia, Qiantong Xu, Naman Goyal, Kritika Singh, Patrick von Platen, Yatharth Saraf, Juan Pino, Alexei Baevski, Alexis Conneau, Michael Auli

Extracting Targeted Training Data from ASR Models, and How to Mitigate It
Ehsan Amid, Om Thakkar, Arun Narayanan, Rajiv Mathews, Françoise Beaufays

Detecting Unintended Memorization in Language-Model-Fused ASR
W. Ronny Huang, Steve Chien, Om Thakkar, Rajiv Mathews

AVATAR: Unconstrained Audiovisual Speech Recognition
Valentin Gabeur, Paul Hongsuck Seo, Arsha Nagrani, Chen Sun, Karteek Alahari, Cordelia Schmid

End-to-End Multi-talker Audio-Visual ASR Using an Active Speaker Attention Module
Richard Rose, Olivier Siohan

Transformer-Based Video Front-Ends for Audio-Visual Speech Recognition for Single and Multi-person Video
Dmitriy Serdyuk, Otavio Braga, Olivier Siohan

Unsupervised Data Selection via Discrete Speech Representation for ASR
Zhiyun Lu, Yongqiang Wang, Yu Zhang, Wei Han, Zhehuai Chen, Parisa Haghani

Non-parallel Voice Conversion for ASR Augmentation
Gary Wang, Andrew Rosenberg, Bhuvana Ramabhadran, Fadi Biadsy, Jesse Emond, Yinghui Huang, Pedro J. Moreno

Ultra-Low-Bitrate Speech Coding with Pre-trained Transformers
Ali Siahkoohi, Michael Chinen, Tom Denton, W. Bastiaan Kleijn, Jan Skoglund

Streaming End-to-End Multilingual Speech Recognition with Joint Language Identification
Chao Zhang, Bo Li, Tara Sainath, Trevor Strohman, Sepand Mavandadi, Shuo-Yiin Chang, Parisa Haghani

Improving Deliberation by Text-Only and Semi-supervised Training
Ke Hu, Tara N. Sainath, Yanzhang He, Rohit Prabhavalkar, Trevor Strohman, Sepand Mavandadi, Weiran Wang

E2E Segmenter: Joint Segmenting and Decoding for Long-Form ASR
W. Ronny Huang, Shuo-yiin Chang, David Rybach, Rohit Prabhavalkar, Tara N. Sainath, Cyril Allauzen, Cal Peyser, Zhiyun Lu

CycleGAN-Based Unpaired Speech Dereverberation
Alexis Conneau, Ankur Bapna, Yu Zhang, Min Ma, Patrick von Platen, Anton Lozhkov, Colin Cherry, Ye Jia, Clara Rivera, Mihir Kale, Daan van Esch, Vera Axelrod, Simran Khanuja, Jonathan Clark, Orhan Firat, Michael Auli, Sebastian Ruder, Jason Riesa, Melvin Johnson

TRILLsson: Distilled Universal Paralinguistic Speech Representations (see blog post)
Joel Shor, Subhashini Venugopalan

Learning Neural Audio Features Without Supervision
Sarthak Yadav, Neil Zeghidour

SpeechPainter: Text-Conditioned Speech Inpainting
Zalan Borsos, Matthew Sharifi, Marco Tagliasacchi

SpecGrad: Diffusion Probabilistic Model-Based Neural Vocoder with Adaptive Noise Spectral Shaping
Yuma Koizumi, Heiga Zen, Kohei Yatabe, Nanxin Chen, Michiel Bacchiani

Distance-Based Sound Separation
Katharine Patterson, Kevin Wilson, Scott Wisdom, John R. Hershey

Analysis of Self-Attention Head Diversity for Conformer-Based Automatic Speech Recognition
Kartik Audhkhasi, Yinghui Huang, Bhuvana Ramabhadran, Pedro J. Moreno

Improving Rare Word Recognition with LM-Aware MWER Training
Wang Weiran, Tongzhou Chen, Tara Sainath, Ehsan Variani, Rohit Prabhavalkar, W. Ronny Huang, Bhuvana Ramabhadran, Neeraj Gaur, Sepand Mavandadi, Cal Peyser, Trevor Strohman, Yanzhang He, David Rybach

MAESTRO: Matched Speech Text Representations Through Modality Matching
Zhehuai Chen, Yu Zhang, Andrew Rosenberg, Bhuvana Ramabhadran, Pedro J. Moreno, Ankur Bapna, Heiga Zen

Pseudo Label is Better Than Human Label
Dongseong Hwang, Khe Chai Sim, Zhouyuan Huo, Trevor Strohman

On the Optimal Interpolation Weights for Hybrid Autoregressive Transducer Model
Ehsan Variani, Michael Riley, David Rybach, Cyril Allauzen, Tongzhou Chen, Bhuvana Ramabhadran

Streaming Align-Refine for Non-autoregressive Deliberation
Wang Weiran, Ke Hu, Tara Sainath

Federated Pruning: Improving Neural Network Efficiency with Federated Learning
Rongmei Lin*, Yonghui Xiao, Tien-Ju Yang, Ding Zhao, Li Xiong, Giovanni Motta, Françoise Beaufays

A Unified Cascaded Encoder ASR Model for Dynamic Model Sizes
Shaojin Ding, Weiran Wang, Ding Zhao, Tara N Sainath, Yanzhang He, Robert David, Rami Botros, Xin Wang, Rina Panigrahy, Qiao Liang, Dongseong Hwang, Ian McGraw, Rohit Prabhavalkar, Trevor Strohman

4-Bit Conformer with Native Quantization Aware Training for Speech Recognition
Shaojin Ding, Phoenix Meadowlark, Yanzhang He, Lukasz Lew, Shivani Agrawal, Oleg Rybakov

Visually-Aware Acoustic Event Detection Using Heterogeneous Graphs
Amir Shirian, Krishna Somandepalli, Victor Sanchez, Tanaya Guha

A Conformer-Based Waveform-Domain Neural Acoustic Echo Canceller Optimized for ASR Accuracy
Sankaran Panchapagesan, Arun Narayanan, Turaj Zakizadeh Shabestary, Shuai Shao, Nathan Howard, Alex Park, James Walker, Alexander Gruenstein

Reducing Domain Mismatch in Self-Supervised Speech Pre-training
Murali Karthick Baskar, Andrew Rosenberg, Bhuvana Ramabhadran, Yu Zhang, Nicolás Serrano

On-the-Fly ASR Corrections with Audio Exemplars
Golan Pundak, Tsendsuren Munkhdalai, Khe Chai Sim

A Language Agnostic Multilingual Streaming On-Device ASR System
Bo Li, Tara Sainath, Ruoming Pang*, Shuo-Yiin Chang, Qiumin Xu, Trevor Strohman, Vince Chen, Qiao Liang, Heguang Liu, Yanzhang He, Parisa Haghani, Sameer Bidichandani

XTREME-S: Evaluating Cross-Lingual Speech Representations
Alexis Conneau, Ankur Bapna, Yu Zhang, Min Ma, Patrick von Platen, Anton Lozhkov, Colin Cherry, Ye Jia, Clara Rivera, Mihir Kale, Daan van Esch, Vera Axelrod, Simran Khanuja, Jonathan Clark, Orhan Firat, Michael Auli, Sebastian Ruder, Jason Riesa, Melvin Johnson

Towards Disentangled Speech Representations
Cal Peyser, Ronny Huang, Andrew Rosenberg, Tara Sainath, Michael Picheny, Kyunghyun Cho

Personal VAD 2.0: Optimizing Personal Voice Activity Detection for On-Device Speech Recognition
Shaojin Ding, Rajeev Rikhye, Qiao Liang, Yanzhang He, Quan Wang, Arun Narayanan, Tom O'Malley, Ian McGraw

A Universally-Deployable ASR Frontend for Joint Acoustic Echo Cancellation, Speech Enhancement, and Voice Separation
Tom O’Malley, Arun Narayanan, Quan Wang

Training Text-To-Speech Systems From Synthetic Data: A Practical Approach For Accent Transfer Tasks
Lev Finkelstein, Heiga Zen, Norman Casagrande, Chun-an Chan, Ye Jia, Tom Kenter, Alex Petelin, Jonathan Shen*, Vincent Wan, Yu Zhang, Yonghui Wu, Robert Clark

A Scalable Model Specialization Framework for Training and Inference Using Submodels and Its Application to Speech Model Personalization
Fadi Biadsy, Youzheng Chen, Xia Zhang, Oleg Rybakov, Andrew Rosenberg, Pedro Moreno

Text-Driven Separation of Arbitrary Sounds
Kevin Kilgour, Beat Gfeller, Qingqing Huang, Aren Jansen, Scott Wisdom, Marco Tagliasacchi


Workshops, Tutorials & Special Sessions

The VoxCeleb Speaker Recognition Challenge 2022 (VoxSRC-22)
Organizers include: Arsha Nagrani

Self-Supervised Representation Learning for Speech Processing
Organizers include: Tara Sainath

Learning from Weak Labels
Organizers include: Ankit Shah

RNN Transducers for Named Entity Recognition with Constraints on Alignment for Understanding Medical Conversations
Authors: Hagen Soltau, Izhak Shafran, Mingqiu Wang, Laurent El Shafey

Listening with Googlears: Low-Latency Neural Multiframe Beamforming and Equalization for Hearing Aids
Authors: Samuel Yang, Scott Wisdom, Chet Gnegy, Richard F. Lyon, Sagar Savla

Using Rater and System Metadata to Explain Variance in the VoiceMOS Challenge 2022 Dataset
Authors: Michael Chinen, Jan Skoglund, Chandan K. A. Reddy, Alessandro Ragano, Andrew Hines

Incremental Layer-Wise Self-Supervised Learning for Efficient Unsupervised Speech Domain Adaptation On Device
Authors: Zhouyuan Huo, Dongseong Hwang, Khe Chai Sim, Shefali Garg, Ananya Misra, Nikhil Siddhartha, Trevor Strohman, Françoise Beaufays

Trustworthy Speech Processing
Organizers include: Shrikanth Narayanan



*Work done while at Google.  

Source: Google AI Blog


Introducing CVSS: A Massively Multilingual Speech-to-Speech Translation Corpus

Automatic translation of speech from one language to speech in another language, called speech-to-speech translation (S2ST), is important for breaking down the communication barriers between people speaking different languages. Conventionally, automatic S2ST systems are built with a cascade of automatic speech recognition (ASR), text-to-text machine translation (MT), and text-to-speech (TTS) synthesis sub-systems, so that the system overall is text-centric. Recently, work on S2ST that doesn’t rely on intermediate text representation is emerging, such as end-to-end direct S2ST (e.g., Translatotron) and cascade S2ST based on learned discrete representations of speech (e.g., Tjandra et al.). While early versions of such direct S2ST systems obtained lower translation quality compared to cascade S2ST models, they are gaining traction as they have the potential both to reduce translation latency and compounding errors, and to better preserve paralinguistic and non-linguistic information from the original speech, such as voice, emotion, tone, etc. However, such models usually have to be trained on datasets with paired S2ST data, but the public availability of such corpora is extremely limited.

To foster research on such a new generation of S2ST, we introduce a Common Voice-based Speech-to-Speech translation corpus, or CVSS, which includes sentence-level speech-to-speech translation pairs from 21 languages into English. Unlike existing public corpora, CVSS can be directly used for training such direct S2ST models without any extra processing. In “CVSS Corpus and Massively Multilingual Speech-to-Speech Translation”, we describe the dataset design and development, and demonstrate the effectiveness of the corpus through training of baseline direct and cascade S2ST models and showing performance of a direct S2ST model that approaches that of a cascade S2ST model.

Building CVSS
CVSS is directly derived from the CoVoST 2 speech-to-text (ST) translation corpus, which is further derived from the Common Voice speech corpus. Common Voice is a massively multilingual transcribed speech corpus designed for ASR in which the speech is collected by contributors reading text content from Wikipedia and other text corpora. CoVoST 2 further provides professional text translation for the original transcript from 21 languages into English and from English into 15 languages. CVSS builds on these efforts by providing sentence-level parallel speech-to-speech translation pairs from 21 languages into English (shown in the table below).

To facilitate research with different focuses, two versions of translation speech in English are provided in CVSS, both are synthesized using state-of-the-art TTS systems, with each version providing unique value that doesn’t exist in other public S2ST corpora:

  • CVSS-C: All the translation speech is in a single canonical speaker’s voice. Despite being synthetic, the speech is highly natural, clean, and consistent in speaking style. These properties ease the modeling of the target speech and enable trained models to produce high quality translation speech suitable for general user-facing applications where speech quality is of higher importance than accurately reproducing the speakers' voices.
  • CVSS-T: The translation speech captures the voice from the corresponding source speech. Each S2ST pair has a similar voice on the two sides, despite being in different languages. Because of this, the dataset is suitable for building models where accurate voice preservation is desired, such as for movie dubbing.

Together with the source speech, the two S2ST datasets contain 1,872 and 1,937 hours of speech, respectively.

Source
Language    
Code     Source
  speech (X)  
CVSS-C
  target speech (En)  
CVSS-T
  target speech (En)  
French fr 309.3 200.3 222.3
German de 226.5 137.0 151.2
Catalan ca 174.8 112.1 120.9
Spanish es 157.6 94.3 100.2
Italian it 73.9 46.5 49.2
Persian fa 58.8 29.9 34.5
Russian ru 38.7 26.9 27.4
Chinese zh 26.5 20.5 22.1
Portuguese     pt 20.0 10.4 11.8
Dutch nl 11.2 7.3 7.7
Estonian et 9.0 7.3 7.1
Mongolian mn 8.4 5.1 5.7
Turkish tr 7.9 5.4 5.7
Arabic ar 5.8 2.7 3.1
Latvian lv 4.9 2.6 3.1
Swedish sv 4.3 2.3 2.8
Welsh cy 3.6 1.9 2.0
Tamil ta 3.1 1.7 2.0
Indonesian id 3.0 1.6 1.7
Japanese ja 3.0 1.7 1.8
Slovenian sl 2.9 1.6 1.9
Total 1,153.2 719.1 784.2
Amount of source and target speech of each X-En pair in CVSS (hours).

In addition to translation speech, CVSS also provides normalized translation text matching the pronunciation in the translation speech (on numbers, currencies, acronyms, etc., see data samples below, e.g., where “100%” is normalized as “one hundred percent” or “King George II” is normalized as “king george the second”), which can benefit both model training as well as standardizing the evaluation.

CVSS is released under the Creative Commons Attribution 4.0 International (CC BY 4.0) license and it can be freely downloaded online.

Data Samples

Example 1:
Source audio (French)   
Source transcript (French)    Le genre musical de la chanson est entièrement le disco.
CVSS-C translation audio (English)   
CVSS-T translation audio (English)   
Translation text (English)    The musical genre of the song is 100% Disco.
Normalized translation text (English)        the musical genre of the song is one hundred percent disco
   
   
Example 2:
Source audio (Chinese)       
Source transcript (Chinese)        弗雷德里克王子,英国王室成员,为乔治二世之孙,乔治三世之幼弟。
CVSS-C translation audio (English)       
CVSS-T translation audio (English)       
Translation text (English)        Prince Frederick, member of British Royal Family, Grandson of King George II, brother of King George III.
Normalized translation text (English)        prince frederick member of british royal family grandson of king george the second brother of king george the third

Baseline Models
On each version of CVSS, we trained a baseline cascade S2ST model as well as two baseline direct S2ST models and compared their performance. These baselines can be used for comparison in future research.

Cascade S2ST: To build strong cascade S2ST baselines, we trained an ST model on CoVoST 2, which outperforms the previous states of the art by +5.8 average BLEU on all 21 language pairs (detailed in the paper) when trained on the corpus without using extra data. This ST model is connected to the same TTS models used for constructing CVSS to compose very strong cascade S2ST baselines (ST → TTS).

Direct S2ST: We built two baseline direct S2ST models using Translatotron and Translatotron 2. When trained from scratch with CVSS, the translation quality from Translatotron 2 (8.7 BLEU) approaches that of the strong cascade S2ST baseline (10.6 BLEU). Moreover, when both use pre-training the gap decreases to only 0.7 BLEU on ASR transcribed translation. These results verify the effectiveness of using CVSS to train direct S2ST models.

Translation quality of baseline direct and cascade S2ST models built on CVSS-C, measured by BLEU on ASR transcription from speech translation. The pre-training was done on CoVoST 2 without other extra data sets.

Conclusion
We have released two versions of multilingual-to-English S2ST datasets, CVSS-C and CVSS-T, each with about 1.9K hours of sentence-level parallel S2ST pairs, covering 21 source languages. The translation speech in CVSS-C is in a single canonical speaker’s voice, while the same in CVSS-T is in voices transferred from the source speech. Each of these datasets provides unique value not existing in other public S2ST corpora.

We built baseline multilingual direct S2ST models and cascade S2ST models on both datasets, which can be used for comparison in future works. To build strong cascade S2ST baselines, we trained an ST model on CoVoST 2, which outperforms the previous states of the art by +5.8 average BLEU when trained on the corpus without extra data. Nevertheless, the performance of the direct S2ST models approaches the strong cascade baselines when trained from scratch, and with only 0.7 BLEU difference on ASR transcribed translation when utilized pre-training. We hope this work helps accelerate the research on direct S2ST.

Acknowledgments
We acknowledge the volunteer contributors and the organizers of the Common Voice and LibriVox projects for their contribution and collection of recordings, the creators of Common Voice, CoVoST, CoVoST 2, Librispeech and LibriTTS corpora for their previous work. The direct contributors to the CVSS corpus and the paper include Ye Jia, Michelle Tadmor Ramanovich, Quan Wang, Heiga Zen. We also thank Ankur Bapna, Yiling Huang, Jason Pelecanos, Colin Cherry, Alexis Conneau, Yonghui Wu, Hadar Shemtov and Françoise Beaufays for helpful discussions and support.

Source: Google AI Blog


Introducing CVSS: A Massively Multilingual Speech-to-Speech Translation Corpus

Automatic translation of speech from one language to speech in another language, called speech-to-speech translation (S2ST), is important for breaking down the communication barriers between people speaking different languages. Conventionally, automatic S2ST systems are built with a cascade of automatic speech recognition (ASR), text-to-text machine translation (MT), and text-to-speech (TTS) synthesis sub-systems, so that the system overall is text-centric. Recently, work on S2ST that doesn’t rely on intermediate text representation is emerging, such as end-to-end direct S2ST (e.g., Translatotron) and cascade S2ST based on learned discrete representations of speech (e.g., Tjandra et al.). While early versions of such direct S2ST systems obtained lower translation quality compared to cascade S2ST models, they are gaining traction as they have the potential both to reduce translation latency and compounding errors, and to better preserve paralinguistic and non-linguistic information from the original speech, such as voice, emotion, tone, etc. However, such models usually have to be trained on datasets with paired S2ST data, but the public availability of such corpora is extremely limited.

To foster research on such a new generation of S2ST, we introduce a Common Voice-based Speech-to-Speech translation corpus, or CVSS, which includes sentence-level speech-to-speech translation pairs from 21 languages into English. Unlike existing public corpora, CVSS can be directly used for training such direct S2ST models without any extra processing. In “CVSS Corpus and Massively Multilingual Speech-to-Speech Translation”, we describe the dataset design and development, and demonstrate the effectiveness of the corpus through training of baseline direct and cascade S2ST models and showing performance of a direct S2ST model that approaches that of a cascade S2ST model.

Building CVSS
CVSS is directly derived from the CoVoST 2 speech-to-text (ST) translation corpus, which is further derived from the Common Voice speech corpus. Common Voice is a massively multilingual transcribed speech corpus designed for ASR in which the speech is collected by contributors reading text content from Wikipedia and other text corpora. CoVoST 2 further provides professional text translation for the original transcript from 21 languages into English and from English into 15 languages. CVSS builds on these efforts by providing sentence-level parallel speech-to-speech translation pairs from 21 languages into English (shown in the table below).

To facilitate research with different focuses, two versions of translation speech in English are provided in CVSS, both are synthesized using state-of-the-art TTS systems, with each version providing unique value that doesn’t exist in other public S2ST corpora:

  • CVSS-C: All the translation speech is in a single canonical speaker’s voice. Despite being synthetic, the speech is highly natural, clean, and consistent in speaking style. These properties ease the modeling of the target speech and enable trained models to produce high quality translation speech suitable for general user-facing applications where speech quality is of higher importance than accurately reproducing the speakers' voices.
  • CVSS-T: The translation speech captures the voice from the corresponding source speech. Each S2ST pair has a similar voice on the two sides, despite being in different languages. Because of this, the dataset is suitable for building models where accurate voice preservation is desired, such as for movie dubbing.

Together with the source speech, the two S2ST datasets contain 1,872 and 1,937 hours of speech, respectively.

Source
Language    
Code     Source
  speech (X)  
CVSS-C
  target speech (En)  
CVSS-T
  target speech (En)  
French fr 309.3 200.3 222.3
German de 226.5 137.0 151.2
Catalan ca 174.8 112.1 120.9
Spanish es 157.6 94.3 100.2
Italian it 73.9 46.5 49.2
Persian fa 58.8 29.9 34.5
Russian ru 38.7 26.9 27.4
Chinese zh 26.5 20.5 22.1
Portuguese     pt 20.0 10.4 11.8
Dutch nl 11.2 7.3 7.7
Estonian et 9.0 7.3 7.1
Mongolian mn 8.4 5.1 5.7
Turkish tr 7.9 5.4 5.7
Arabic ar 5.8 2.7 3.1
Latvian lv 4.9 2.6 3.1
Swedish sv 4.3 2.3 2.8
Welsh cy 3.6 1.9 2.0
Tamil ta 3.1 1.7 2.0
Indonesian id 3.0 1.6 1.7
Japanese ja 3.0 1.7 1.8
Slovenian sl 2.9 1.6 1.9
Total 1,153.2 719.1 784.2
Amount of source and target speech of each X-En pair in CVSS (hours).

In addition to translation speech, CVSS also provides normalized translation text matching the pronunciation in the translation speech (on numbers, currencies, acronyms, etc., see data samples below, e.g., where “100%” is normalized as “one hundred percent” or “King George II” is normalized as “king george the second”), which can benefit both model training as well as standardizing the evaluation.

CVSS is released under the Creative Commons Attribution 4.0 International (CC BY 4.0) license and it can be freely downloaded online.

Data Samples

Example 1:
Source audio (French)   
Source transcript (French)    Le genre musical de la chanson est entièrement le disco.
CVSS-C translation audio (English)   
CVSS-T translation audio (English)   
Translation text (English)    The musical genre of the song is 100% Disco.
Normalized translation text (English)        the musical genre of the song is one hundred percent disco
   
   
Example 2:
Source audio (Chinese)       
Source transcript (Chinese)        弗雷德里克王子,英国王室成员,为乔治二世之孙,乔治三世之幼弟。
CVSS-C translation audio (English)       
CVSS-T translation audio (English)       
Translation text (English)        Prince Frederick, member of British Royal Family, Grandson of King George II, brother of King George III.
Normalized translation text (English)        prince frederick member of british royal family grandson of king george the second brother of king george the third

Baseline Models
On each version of CVSS, we trained a baseline cascade S2ST model as well as two baseline direct S2ST models and compared their performance. These baselines can be used for comparison in future research.

Cascade S2ST: To build strong cascade S2ST baselines, we trained an ST model on CoVoST 2, which outperforms the previous states of the art by +5.8 average BLEU on all 21 language pairs (detailed in the paper) when trained on the corpus without using extra data. This ST model is connected to the same TTS models used for constructing CVSS to compose very strong cascade S2ST baselines (ST → TTS).

Direct S2ST: We built two baseline direct S2ST models using Translatotron and Translatotron 2. When trained from scratch with CVSS, the translation quality from Translatotron 2 (8.7 BLEU) approaches that of the strong cascade S2ST baseline (10.6 BLEU). Moreover, when both use pre-training the gap decreases to only 0.7 BLEU on ASR transcribed translation. These results verify the effectiveness of using CVSS to train direct S2ST models.

Translation quality of baseline direct and cascade S2ST models built on CVSS-C, measured by BLEU on ASR transcription from speech translation. The pre-training was done on CoVoST 2 without other extra data sets.

Conclusion
We have released two versions of multilingual-to-English S2ST datasets, CVSS-C and CVSS-T, each with about 1.9K hours of sentence-level parallel S2ST pairs, covering 21 source languages. The translation speech in CVSS-C is in a single canonical speaker’s voice, while the same in CVSS-T is in voices transferred from the source speech. Each of these datasets provides unique value not existing in other public S2ST corpora.

We built baseline multilingual direct S2ST models and cascade S2ST models on both datasets, which can be used for comparison in future works. To build strong cascade S2ST baselines, we trained an ST model on CoVoST 2, which outperforms the previous states of the art by +5.8 average BLEU when trained on the corpus without extra data. Nevertheless, the performance of the direct S2ST models approaches the strong cascade baselines when trained from scratch, and with only 0.7 BLEU difference on ASR transcribed translation when utilized pre-training. We hope this work helps accelerate the research on direct S2ST.

Acknowledgments
We acknowledge the volunteer contributors and the organizers of the Common Voice and LibriVox projects for their contribution and collection of recordings, the creators of Common Voice, CoVoST, CoVoST 2, Librispeech and LibriTTS corpora for their previous work. The direct contributors to the CVSS corpus and the paper include Ye Jia, Michelle Tadmor Ramanovich, Quan Wang, Heiga Zen. We also thank Ankur Bapna, Yiling Huang, Jason Pelecanos, Colin Cherry, Alexis Conneau, Yonghui Wu, Hadar Shemtov and Françoise Beaufays for helpful discussions and support.

Source: Google AI Blog


TRILLsson: Small, Universal Speech Representations for Paralinguistic Tasks

In recent years, we have seen dramatic improvements on lexical tasks such as automatic speech recognition (ASR). However, machine systems still struggle to understand paralinguistic aspects — such as tone, emotion, whether a speaker is wearing a mask, etc. Understanding these aspects represents one of the remaining difficult problems in machine hearing. In addition, state-of-the-art results often come from ultra-large models trained on private data, making them impractical to run on mobile devices or to release publicly.

In “Universal Paralinguistic Speech Representations Using Self-Supervised Conformers”, to appear in ICASSP 2022, we introduce CAP12— the 12th layer of a 600M parameter model trained on the YT-U training dataset using self-supervision. We demonstrate that the CAP12 model outperforms nearly all previous results in our paralinguistic benchmark, sometimes by large margins, even though previous results are often task-specific. In “TRILLsson: Distilled Universal Paralinguistic Speech Representations'', we introduce the small, performant, publicly-available TRILLsson models and demonstrate how we reduced the size of the high-performing CAP12 model by 6x-100x while maintaining 90-96% of the performance. To create TRILLsson, we apply knowledge distillation on appropriately-sized audio chunks and use different architecture types to train smaller, faster networks that are small enough to run on mobile devices.

1M-Hour Dataset to Train Ultra-Large Self-Supervised Models
We leverage the YT-U training dataset to train the ultra-large, self-supervised CAP12 model. The YT-U dataset is a highly varied, 900M+ hour dataset that contains audio of various topics, background conditions, and speaker acoustic properties.

Video categories by length (outer) and number (inner), demonstrating the variety in the YT-U dataset (figure from BigSSL)

We then modify a Wav2Vec 2.0 self-supervised training paradigm, which can solve tasks using raw data without labels, and combine it with ultra-large Conformer models. Because self-training doesn't require labels, we can take full advantage of YT-U by scaling up our models to some of the largest model sizes ever trained, including 600M, 1B, and 8B parameters.

NOSS: A Benchmark for Paralinguistic Tasks
We demonstrate that an intermediate representation of one of the previous models contains a state-of-the-art representation for paralinguistic speech. We call the 600M parameter Conformer model without relative attention Conformer Applied to Paralinguistics (CAP). We exhaustively search through all intermediate representations of six ultra-large models and find that layer 12 (CAP12) outperforms previous representations by significant margins.

To measure the quality of the roughly 300 candidate paralinguistic speech representations, we evaluate on an expanded version of the NOn-Semantic Speech (NOSS) benchmark, which is a collection of well-studied paralinguistic speech tasks, such as speech emotion recognition, language identification, and speaker identification. These tasks focus on paralinguistics aspects of speech, which require evaluating speech features on the order of 1 second or longer, rather than lexical features, which require 100ms or shorter. We then add to the benchmark a mask-wearing task introduced at Interspeech 2020, a fake speech detection task (ASVSpoof 2019), a task to detect the level of dysarthria from project Euphonia, and an additional speech emotion recognition task (IEMOCAP). By expanding the benchmark and increasing the diversity of the tasks, we empirically demonstrate that CAP12 is even more generally useful than previous representations.

Simple linear models on time-averaged CAP12 representations even outperform complex, task-specific models on five out of eight paralinguistic tasks. This is surprising because comparable models sometimes use additional modalities (e.g., vision and speech, or text and speech) as well. Furthermore, CAP12 is exceptionally good at emotion recognition tasks. CAP12 embeddings also outperform all other embeddings on all other tasks with only a single exception: for one embedding from a supervised network on the dysarthria detection task.

Model Voxceleb   Voxforge   Speech Commands   ASVSpoof2019∗∗   Euphonia#   CREMA-D   IEMOCAP
Prev SoTA - 95.4 97.9 5.11 45.9 74.0 67.6+
TRILL 12.6 84.5 77.6 74.6 48.1 65.7 54.3
ASR Embedding 5.2 98.9 96.1 11.2 54.5 71.8 65.4
Wav2Vec2 layer 6†† 17.9 98.5 95.0 6.7 48.2 77.4 65.8
CAP12 51.0 99.7 97.0 2.5 51.5 88.2 75.0
Test performance on the NOSS Benchmark and extended tasks. “Prev SoTA” indicates the previous best performing state-of-the-art model, which has arbitrary complexity, but all other rows are linear models on time-averaged input. Filtered according to YouTube’s privacy guidelines. ∗∗ Uses equal error rate [20]. # The only non-public dataset. We exclude it from aggregate scores. Audio and visual features used in previous state-of-the-art models. + The previous state-of-the-art model performed cross-validation. For our evaluation, we hold out two specific speakers as a test. †† Wav2Vec 2.0 model from HuggingFace. Best overall layer was layer 6.

TRILLsson: Small, High Quality, Publicly Available Models
Similar to FRILL, our next step was to make an on-device, publicly available version of CAP12. This involved using knowledge distillation to train smaller, faster, mobile-friendly architectures. We experimented with EfficientNet, Audio Spectrogram Transformer (AST), and ResNet. These model types are very different, and cover both fixed-length and arbitrary-length inputs. EfficientNet comes from a neural architecture search over vision models to find simultaneously performant and efficient model structures. AST models are transformers adapted to audio inputs. ResNet is a standard architecture that has shown good performance across many different models.

We trained models that performed on average 90-96% as well as CAP12, despite being 1%-15% the size and trained using only 6% the data. Interestingly, we found that different architecture types performed better at different sizes. ResNet models performed best at the low end, EfficientNet in the middle, and AST models at the larger end.

Aggregate embedding performance vs. model size for various student model architectures and sizes. We demonstrate that ResNet architectures perform best for small sizes, EfficientNetV2 performs best in the midsize model range, up to the largest model size tested, after which the larger AST models are best.

We perform knowledge distillation with the goal of matching a student, with a fixed-size input, to the output of a teacher, with a variable-size input, for which there are two methods of generating student targets: global matching and local matching. Global matching produces distillation targets by generating CAP12 embeddings for an entire audio clip, and then requires that a student match the target from just a small segment of audio (e.g., 2 seconds). Local matching requires that the student network match the average CAP12 embedding just over the smaller portion of the audio that the student sees. In our work, we focused on local matching.

Two types of generating distillation targets for sequences. Left: Global matching uses the average CAP12 embedding over the whole clip for the target for each local chunk. Right: Local matching uses CAP12 embeddings averaged just over local clips as the distillation target.

Observation of Bimodality and Future Directions
Paralinguistic information shows an unexpected bimodal distribution. For the CAP model that operates on 500 ms input segments, and two of the full-input Conformer models, intermediate representations gradually increase in paralinguistic information, then decrease, then increase again, and finally lose this information towards the output layer. Surprisingly, this pattern is also seen when exploring the intermediate representations of networks trained on retinal images.

500 ms inputs to CAP show a relatively pronounced bimodal distribution of paralinguistic information across layers.
Two of the conformer models with full inputs show a bimodal distribution of paralinguistic information across layers.

We hope that smaller, faster models for paralinguistic speech unlock new applications in speech recognition, text-to-speech generation, and understanding user intent. We also expect that smaller models will be more easily interpretable, which will allow researchers to understand what aspects of speech are important for paralinguistics. Finally, we hope that our open-sourced speech representations are used by the community to improve paralinguistic speech tasks and user understanding in private or small datasets.

Acknowledgements
I'd like to thank my co-authors Aren Jansen, Wei Han, Daniel Park, Yu Zhang, and Subhashini Venugopalan for their hard work and creativity on this project. I'd also like to thank the members of the large collaboration for the BigSSL work, without which these projects would not be possible. The team includes James Qin, Anmol Gulati, Yuanzhong Xu, Yanping Huang, Shibo Wang, Zongwei Zhou, Bo Li, Min Ma, William Chan, Jiahui Yu, Yongqiang Wang, Liangliang Cao, Khe Chai Sim, Bhuvana Ramabhadran, Tara N. Sainath, Françoise Beaufays, Zhifeng Chen, Quoc V. Le, Chung-Cheng Chiu, Ruoming Pang, and Yonghui Wu.

Source: Google AI Blog